hep-lat0501020/a2.tex
1: \documentclass[aps,prd,preprint,tightenlines,groupedaddress,nofootinbib,showpacs,byrevtex]{revtex4}
2: \usepackage{amssymb,latexsym}
3: \usepackage{amsmath,amsbsy}
4: \usepackage{epsfig,bm}
5: \usepackage{graphicx,comment}
6: \unitlength=1mm
7: \DeclareMathOperator{\st}{str}
8: \DeclareMathOperator{\tr}{tr}
9: 
10: 
11: \begin{document}
12: \def\a{{\alpha}}
13: \def\b{{\beta}}
14: \def\d{{\delta}}
15: \def\D{{\Delta}}
16: \def\e{{\varepsilon}}
17: \def\g{{\gamma}}
18: \def\G{{\Gamma}}
19: \def\k{{\kappa}}
20: \def\l{{\lambda}}
21: \def\L{{\Lambda}}
22: \def\m{{\mu}}
23: \def\n{{\nu}}
24: \def\o{{\omega}}
25: \def\O{{\Omega}}
26: \def\S{{\Sigma}}
27: \def\s{{\sigma}}
28: \def\th{{\theta}}
29: 
30: \def\ol#1{{\overline{#1}}}
31: 
32: 
33: \def\Dslash{D\hskip-0.65em /}
34: 
35: \def\CPT{{$\chi$PT}}
36: \def\QCPT{{Q$\chi$PT}}
37: \def\PQCPT{{PQ$\chi$PT}}
38: \def\tr{\text{tr}}
39: \def\str{\text{str}}
40: \def\diag{\text{diag}}
41: \def\order{{\mathcal O}}
42: 
43: 
44: \def\cC{{\mathcal C}}
45: \def\cB{{\mathcal B}}
46: \def\cT{{\mathcal T}}
47: \def\cQ{{\mathcal Q}}
48: \def\cL{{\mathcal L}}
49: \def\cO{{\mathcal O}}
50: \def\cA{{\mathcal A}}
51: \def\cQ{{\mathcal Q}}
52: \def\cR{{\mathcal R}}
53: \def\cH{{\mathcal H}}
54: \def\cW{{\mathcal W}}
55: \def\cM{{\mathcal M}}
56: \def\cJ{{\mathcal J}}
57: 
58: 
59: \def\eqref#1{{(\ref{#1})}}
60: 
61: %\preprint{DUKE-TH-04-XXX}
62:  
63: \title{Baryon masses at $\cO(a^2)$ in chiral perturbation theory}
64: \author{ Brian C.~Tiburzi}
65: \email[]{bctiburz@phy.duke.edu}
66: \affiliation{Department of Physics\\
67: Duke University\\
68: P.O.~Box 90305\\
69: Durham, NC 27708-0305}
70: 
71: \date{\today}
72: 
73: 
74: \pacs{12.38.Gc}
75: 
76: \begin{abstract}
77: The chiral Lagrangian for the Symanzik action through $\cO(a^2)$ for baryons is obtained.
78: We consider two flavor unquenched and partially quenched lattice theories, allowing for mixed 
79: actions in the latter. As an application, we calculate masses to $\cO(a^2)$ for the 
80: nucleons and deltas, and investigate the corrections due to the violation of $O(4)$ rotational invariance. 
81: These results are contrasted with those in the meson sector for lattice 
82: simulations using mixed and unmixed actions of Wilson and Ginsparg-Wilson quarks.
83: \end{abstract}
84: 
85: 
86: \maketitle
87: 
88: \section{Introduction}
89: Lattice gauge theory provides first principles calculations of strong 
90: interaction physics, where QCD is non-perturbative, and quarks and gluons are confined 
91: in color-neutral hadronic states. 
92: These calculations, however, are severely limited by available computing power, 
93: necessitating the use of quark masses $m_q$ that are much larger than those in reality. 
94: To make physical predictions, one must extrapolate from the quark masses used on the lattice to those of nature.
95: A model independent tool for this extrapolation is to study QCD at hadronic scales 
96: using its low-energy effective theory, 
97: chiral perturbation theory (\CPT).
98: Because \CPT\ provides a systematic expansion involving $m_q/\L_{QCD}$, 
99: one can understand how QCD observables 
100: behave, in principle, as functions of the quark mass. 
101: To address the quenched and partially quenched approximations employed by
102: lattice calculations, \CPT\ has been extended to quenched chiral 
103: perturbation theory (\QCPT)~% 
104: \cite{Morel:1987xk,Sharpe:1992ft,Bernard:1992mk,Labrenz:1996jy,Sharpe:1996qp}
105: and partially quenched chiral perturbation theory (\PQCPT)~%
106: \cite{Bernard:1994sv,Sharpe:1997by,Golterman:1998st,Sharpe:2000bc,Sharpe:2001fh,Sharpe:2003vy}.
107: 
108: 
109: While lattice calculations are limited to unphysically large quark masses, they are restricted further by
110: two additional parameters: the size of the lattice $L$, that is not considerably larger than the system under investigation; 
111: and the lattice spacing $a$, that is not considerably smaller than the relevant hadronic distance scale.
112: To address the issue of finite lattice spacing,
113: \CPT\ has been extended (following the earlier work of~\cite{Sharpe:1998xm,Lee:1999zx})
114: in the meson sector
115: to $\cO(a)$ for the Wilson action~\cite{Rupak:2002sm}, and for mixed lattice actions~\cite{Bar:2002nr}.
116: Corrections at $\cO(a^2)$ in \CPT\ have been pursued~\cite{Bar:2003mh,Aoki:2003yv}.
117: There have also been parallel developments in addressing lattice spacing artifacts 
118: in staggered \CPT\ for mesons~\cite{Aubin:2003mg,Aubin:2003uc,Sharpe:2004is},
119: and heavy mesons~\cite{Aubin:2004xd}; and in twisted mass 
120: QCD~\cite{Munster:2003ba,Munster:2004dj,Scorzato:2004da,Sharpe:2004ps,Sharpe:2004ny}.
121: Corrections to baryon observables in \CPT\ and \PQCPT, too, have been recently investigated~\cite{Beane:2003xv,Arndt:2004we}. 
122: To consider finite lattice spacing corrections, one must formulate the underlying lattice theory and match
123: the new operators that appear onto those in the chiral effective theory. This can be done by utilizing a dual 
124: expansion in quark mass and lattice spacing. For an overview, see~\cite{Baer:2004xp}. 
125: Following~\cite{Bar:2003mh,Beane:2003xv}, we assume a hierarchy of energy scales
126: \begin{equation}
127: m_q \ll \L_{QCD} \ll \frac{1}{a}
128: .\end{equation}
129: We shall further choose a power counting scheme in which
130: the small dimensionless expansion parameters are\footnote{%
131: In practice the power counting scheme and subsequent ordering of the dual expansion in 
132: quark mass and lattice spacing should be organized based the on the actual sizes of $m_q$ and $a$. 
133: }
134: \begin{equation} \label{eqn:pc}
135: \e^2 \sim 
136: \begin{cases}
137:  m_q/\L_{QCD}, \\
138:  a \, \L_{QCD}
139: \end{cases}
140: .\end{equation}
141: Thus we have a systematic way to calculate $a$-dependent corrections
142: in \CPT\ for the observables of interest. Such expressions allow one to perform the quark mass extrapolation
143: before the continuum extrapolation. 
144: 
145: 
146: 
147: In this work we address the extension of \CPT\ and \PQCPT\ at finite lattice spacing to $\cO(a^2)$ in 
148: the baryon sector. Specifically we detail the operators which must be included to determine
149: the baryon masses to $\cO(\e^4)$ in the above power counting. Unlike the extension of \CPT\ and \PQCPT\ 
150: to $\cO(a^2)$ in the meson sector, the baryon sector suffers from a proliferation of new operators. 
151: Despite this fact, the number of free independent parameters entering expressions for the masses
152: is still relatively small. Moreover in contrast to the meson sector at $\cO(a^2)$, heavy baryon 
153: operators that break the $O(4)$ rotational symmetry of Euclidean space are required at this order in 
154: the chiral expansion. Such operators are required for particles that are heavy compared to $\L_{QCD}$. 
155: 
156: 
157: 
158: This paper has the following organization. First in Sec.~\ref{s:sym}, we review the Symanzik Lagrangian
159: at $\cO(a^2)$ for a partially quenched, mixed lattice action, where the valence and sea quarks
160: are either Wilson or Ginsparg-Wilson fermions. We focus on the symmetries of the Symanzik Lagrangian because
161: these are essential in constructing the effective theory. Next in Sec.~\ref{s:mesons}, we review the finite lattice
162: spacing corrections at $\cO(a)$ in the meson sector of \PQCPT.  Higher-order corrections are not needed for 
163: baryon observables to the order we work. In Sec.~\ref{s:baryons}, we extend heavy baryon \PQCPT\ to $\cO(\e^4)$. 
164: This includes the addition of operators at $\cO(m_Q \, a)$ and $\cO(a^2)$. Applications of this development
165: are pursued in Sec.~\ref{s:mass}, where we obtain the lattice spacing corrections to the masses of nucleons and deltas. 
166: The unquenched two-flavor theory is addressed in Appendix~\ref{s:2}.
167: Corrections from $O(4)$ breaking operators are treated in detail for particles of spin less than two in Appendix~\ref{s:RPI}.
168: A summary (Sec.~\ref{s:summy}) highlights the lattice spacing corrections in the baryon sector. 
169: Here we contrast the number of independent parameters entering expressions at $\cO(\e^4)$ for baryon masses in the various
170: mixed and unmixed partially quenched theories, as well as in the unquenched theory. 
171: 
172: 
173: 
174: 
175: 
176: 
177: \section{\label{sec:PQCPT}\PQCPT\ at $\cO(a^2)$}
178: 
179: 
180: 
181: To extend the baryon chiral Lagrangian to $\cO(a^2)$, we first review 
182: the Symanzik Lagrangian at $\cO(a^2)$ and the construction of \PQCPT\ in the meson sector. 
183: In this and the following sections, we consider a partially quenched theory with a mixed action. 
184: The result for unquenched simulations is contained in Appendix~\ref{s:2}.
185: 
186: 
187: 
188: 
189: 
190: 
191: \subsection{Symanzik Lagrangian}  \label{s:sym}
192: 
193: 
194: The Symanzik action is the continuum effective theory of the lattice 
195: action~\cite{Symanzik:1983dc,Symanzik:1983gh}. As such it is constructed from continuum 
196: operators based on the symmetries of the underlying lattice theory. The Symanzik 
197: Lagrangian is organized in powers of the lattice spacing $a$, namely
198: \begin{equation}
199: \cL = 
200: \cL^{(4)} 
201: + 
202: a \, \cL^{(5)} 
203: + 
204: a^2 \, \cL^{(6)} 
205: + 
206: \dots \label{eq:syman}
207: ,\end{equation}
208: where $\cL^{(n)}$ represents contributions for dimension-$n$ operators.\footnote{%
209: Not all of the $a$-dependence is parametrized in Eq.~\eqref{eq:syman}. The coefficients of 
210: terms in the Lagrangian $\cL^{(n)}$ depend upon the gauge coupling and thus can have a 
211: weak logarithmic dependence on $a$. Such dependence is beyond the scope of this work. 
212: } 
213: In the continuum limit, $a \to 0$, only the dimension-four operators survive.
214: The Symanzik Lagrangian for the Wilson action was discussed to $\cO(a^2)$ 
215: in~\cite{Sheikholeslami:1985ij} and the analysis to $\cO(a)$ was refined by~\cite{Luscher:1996sc}. 
216: Here we consider the general case of a mixed action in partially quenched QCD (PQQCD). Such an action
217: allows the valence and sea quarks to have different masses. Additionally the valence and sea quarks 
218: can be different types of lattice fermions. 
219: The mixed lattice action has the usual parity invariance, charge conjugation invariance and $SU(N_c)$ gauge symmetry. 
220: Accordingly the Symanzik Lagrangian $\cL$ respects these symmetries order-by-order in $a$. 
221: Because spacetime has been discretized, the $O(4)$ rotational symmetry of continuum Euclidean
222: field theory has been reduced to the hypercubic group.
223: 
224: 
225: The flavor symmetry group of the mixed lattice 
226: action $G$ is generally a direct product of flavor symmetry groups for the particular species of 
227: fermion at hand~\cite{Bar:2002nr}. The mixed action contains different forms of the Dirac 
228: operator for each species of fermion, thus there is no symmetry transformation between the 
229: valence and sea sectors.  At zero quark mass, the flavor symmetry group of the mixed lattice action is 
230: \begin{equation} 
231: G = G_{\text{valence}} \otimes G_{\text{sea}}
232: \label{eq:mixsym}
233: ,\end{equation}
234: where for two quark flavors
235: \begin{equation}
236: G_{\text{valence}} = SU(2|2)_L \otimes SU(2|2)_R
237: ,\end{equation}
238: and
239: \begin{equation}
240: G_{\text{sea}} = SU(2)_L \otimes SU(2)_R
241: .\end{equation}
242: For simplicity we have included the ghost quarks in the valence sector. 
243: The quark mass term breaks these respective chiral symmetries in the valence and sea sectors 
244: down to vector symmetries.  
245: 
246: 
247: 
248: For Wilson fermions~\cite{Wilson:1974sk}, the breaking of chiral symmetry persists even at zero quark mass 
249: due to lattice discretization effects. These effects can be tamed to enter at $\cO(a^2)$ by the so-called $\cO(a)$ improvement.  
250: For fermions satisfying the Ginsparg-Wilson relation~\cite{Ginsparg:1982bj}
251: (e.g., Kaplan fermions~\cite{Kaplan:1992bt}, or overlap fermions~\cite{Narayanan:1993ss}), 
252: the chiral symmetry on the lattice at zero quark mass remains exact at finite $a$~\cite{Luscher:1998pq}.
253: For a recent review of chiral symmetry in lattice theories, see~\cite{Chandrasekharan:2004cn}. 
254: We have left unspecified which species lives in which sector of the theory. As a result of this generality, 
255: we are building the Symanzik Lagrangian for the four possible combinations of valence and sea fermions.
256: Part of this generality is an academic pursuit because we do not anticipate lattice calculations employing 
257: Wilson valence quarks in a Ginsparg-Wilson sea, whereas GW valence quarks with dynamical Wilson fermions
258: is a scenario with potential computational benefits. It is likely, however,  that  
259: Wilson quark masses may never reach the chiral regime~\cite{Beane:2004ks} in which case one must deal with alternate methods
260: to modify~\cite{Leinweber:1999ig} or to improve~\cite{Bernard:2003rp} baryon \CPT.
261: These reservations aside, our goal is to contrast the situation at $\cO(a^2)$
262: in the baryon sector with that of the mesons. We will find comparatively that the mesons are rather
263: special with respect to lattice spacing corrections in \CPT.
264: 
265: 
266: Having discussed the symmetries of $\cL$, we now turn to the contributions near the continuum limit.
267: The terms of the $\cL^{(4)}$ Lagrangian are the most familiar and so we discuss them first. 
268: At dimension four, we have the familiar kinetic and Dirac mass terms for the quarks. To be specific, 
269: we work in a partially quenched theory for two light flavors. 
270: The quark part of the leading-order Symanzik Lagrangian reads
271: \begin{equation}\label{eqn:LPQQCD}
272: \cL^{(4)}
273: =
274: \ol Q \, \Dslash \, Q  +  \ol Q \, m_Q \, Q
275: .\end{equation}
276: The six quarks of PQQCD are in the fundamental representation of
277: the graded group $SU(4|2)$%
278: ~\cite{BahaBalantekin:1981kt,BahaBalantekin:1981qy,BahaBalantekin:1982bk}
279: and appear in the vector
280: \begin{equation}
281:   Q=(u,d,j,l,\tilde{u},\tilde{d})^{\text{T}}
282: ,\end{equation}
283: which obeys the graded equal-time commutation relation
284: \begin{equation} \label{eqn:commutation}
285:   Q^\a_i({\bf x}){Q^\b_j}^\dagger({\bf y})
286:   -(-1)^{\eta_i \eta_j}{Q^\b_j}^\dagger({\bf y})Q^\a_i({\bf x})
287:   =
288:   \d^{\a\b}\d_{ij}\d^3({\bf x}-{\bf y})
289: ,\end{equation}
290: where $\a$ and $\b$ are spin, and $i$ and $j$ are flavor indices.
291: The vanishing graded equal-time commutation relations can be written analogously.
292: The grading factor 
293: \begin{equation}
294:    \eta_k
295:    = \left\{ 
296:        \begin{array}{cl}
297:          1 & \text{for } k=1,2,3,4 \\
298:          0 & \text{for } k=5,6
299:        \end{array}
300:      \right.
301: ,\end{equation}
302: incorporates the different statistics of the quarks. 
303: The quark mass matrix in the isospin limit ($m_d = m_u$) of $SU(4|2)$ is given by 
304: \begin{equation}
305:   m_Q=\text{diag}(m_u, m_u, m_j, m_j, m_u, m_u)
306: .\end{equation}
307: In the limit $a \to 0$, and when $m_j=m_u$, one recovers the isospin limit of QCD.
308: The symmetry of $\cL^{(4)}$ in the zero mass limit is $SU(4|2)_L \otimes SU(4|2)_R$,   
309: i.e.~the effects of the mixed action do not show up in the effective theory at leading order. 
310: The mass term in $\cL^{(4)}$ explicitly breaks the graded chiral symmetry down to the graded vector symmetry 
311: $SU(4|2)_V$. 
312: 
313: 
314: 
315: 
316: 
317: 
318: Let us next consider the dimension-five Lagrangian. After field redefinitions, it consists of only the Pauli term given by
319: \begin{equation} \label{eqn:Pauli}
320: \cL^{(5)} 
321: = 
322: c_{SW} \, \ol Q \, \sigma_{\mu \nu} G_{\mu \nu} w_Q \, Q
323: .\end{equation}
324: This term breaks chiral symmetry in precisely the same 
325: way as the quark mass term. 
326: The Sheikholeslami-Wohlert (SW)~\cite{Sheikholeslami:1985ij} 
327: coefficient is $c_{SW}$ and it is accompanied by the Wilson matrix $w_Q$
328: defined by
329: \begin{equation} \label{eqn:sw}
330: w_Q = \text{diag}
331: (w_v,w_v,w_s,w_s,w_v,w_v)
332: .\end{equation}
333: This matrix accounts for 
334: the chiral symmetry properties of the mixed action.  
335: If the quark $Q_i$ is a Wilson fermion, 
336: then $(w_Q)_i = 1$.  Alternately, if $Q_i$ is of the 
337: Ginsparg-Wilson (GW) variety then $(w_Q)_i = 0$. 
338: Since one expects simulations to be performed with 
339: valence quarks that are all of the same species as well as sea quarks 
340: all of the same species, we have labeled the entries in Eq.~\eqref{eqn:sw}
341: by valence ($v$) and sea ($s$) instead of flavor.\footnote{%
342: It is conceivable that three flavor simulations might employ Wilson up and down quarks but with a GW strange
343: quark in order to eliminate the rather large $\cO( a\, m_s)$ corrections. The development in this
344: work for partially quenched theories is general enough to handle this situation by a 
345: modification of Eq.~\eqref{eqn:sw}. I thank W.~Detmold for bringing this to my attention.} 
346: 
347: 
348: To work at $\cO(a^2)$, we must consider the dimension-six Lagrangian. The exact form of $\cL^{(6)}$ 
349: is irrelevant in constructing the chiral effective theory. We need only know which symmetries 
350: are broken and how. We explicitly list terms of the Lagrangian when it is illustrative to do so. 
351: In describing these operators, we introduce the flavor matrix 
352: $\ol w_Q \equiv 1 - w_Q$, which projects onto the GW sector of the theory.  
353: 
354: 
355: 
356: The terms of $\cL^{(6)}$ fall into five classes.
357: The first class of operators consists of  higher-dimensional quark bilinears. 
358: These operators do not break chiral symmetry but have the flavor symmetry 
359: of the mixed action, i.e.~there are distinct operators involving only the valence sector and only the sea sector.
360: Typical examples are $\ol Q \, \Dslash \, {}^3 w_Q \, Q$, and $\ol Q \, \Dslash \, {}^3 \ol w_Q \, Q$. 
361: In class two, there are quark bilinear operators that break chiral symmetry. Simple dimensional counting
362: indicates that there must be an $m_Q$ insertion for these operators to be in $\cL^{(6)}$. Examples include
363: $\ol Q \, m_Q \, D^2 w_Q \, Q$, and $\ol Q \, m_Q \, D^2 \ol w_Q \, Q$. 
364: Class three consists of all four-quark operators that do not break the chiral symmetry
365: of the valence and sea sectors. These class three operators
366: come in three forms due to the different sectors of the theory, e.g.~$(\ol Q \, w_Q \, \gamma_\mu Q)^2$, 
367: $(\ol Q \, \ol w_Q \, \gamma_\mu Q)^2$,  and $ (\ol Q \, w_Q \, \gamma_\mu Q) (\ol Q \, \ol w_Q \, \gamma_\mu Q)$. 
368: Class four operators are four-quark operators that break chiral symmetry. Since there are four fields, chiral 
369: symmetry is broken in a different way than the quark mass term. A typical class four operator is $(\ol Q \, w_Q \, Q)^2$. 
370: Unlike class three, there are no class four operators involving the flavor matrix $\ol w_Q$. 
371: Finally class five operators are those that break $O(4)$ rotation symmetry. There are two such operators 
372: for the mixed lattice action, namely $\ol Q \, w_Q \, \gamma_\mu D_\mu D_\mu D_\mu Q$, and 
373: $\ol Q \, \ol w_Q \, \gamma_\mu D_\mu D_\mu D_\mu Q$.  These classes are summarized in Table~\ref{t:class}. 
374: 
375: 
376: 
377: \begin{table}
378: \caption{Description of the classes of operators in $\cL^{(6)}$. Each representative 
379: class of operator is classified as to whether it is bilinear, chiral symmetry breaking ($\chi$SB),
380: or $O(4)$ breaking. Operators in $\cL^{(6)}$ that are not bilinears are thus four-quark operators. 
381: Additionally the flavor matrices involved for the mixed action are listed.  
382: }
383: %\begin{ruledtabular}
384: \begin{tabular}{c | c c c c c c   }
385: Class & $\qquad$ Bilinear? $\quad$ & $\qquad \chi$SB? $\qquad$ & $O(4)$ breaking? & Flavor Structure  \\
386: \hline
387: $1$ & Yes & No  & No  & $w_Q$, $\ol w_Q$ & \\ 
388: $2$ & Yes & Yes & No  & $m_Q \otimes w_Q$, $m_Q \otimes \ol w_Q$ & \\ 
389: $3$ & No  & No  & No  & $w_Q \otimes w_Q$, $w_Q \otimes \ol w_Q$, $\ol w_Q \otimes \ol w_Q$ & \\ 
390: $4$ & No  & Yes & No  & $w_Q \otimes w_Q$ & \\ 
391: $5$ & Yes & No  & Yes & $w_Q$, $\ol w_Q$ & 
392: \end{tabular}
393: %\end{ruledtabular}
394: \label{t:class}
395: \end{table}
396: 
397: 
398: 
399: 
400: 
401: \subsection{Mesons} \label{s:mesons}
402: 
403: 
404: For massless quarks at zero lattice spacing,
405: the Lagrangian in Eq.~(\ref{eqn:LPQQCD}) exhibits a graded symmetry
406: $SU(4|2)_L \otimes SU(4|2)_R \otimes U(1)_V$ that is assumed 
407: to be spontaneously broken to $SU(4|2)_V \otimes U(1)_V$. 
408: The low-energy effective theory of PQQCD that results from 
409: perturbing about the physical vacuum is \PQCPT.
410: The pseudo-Goldstone mesons can be described at $\cO(\e^2)$ by a Lagrangian 
411: that accounts for the two sources of explicit chiral symmetry breaking:
412: the quark mass term in Eq.~\eqref{eqn:LPQQCD}, and the Pauli term in Eq.~\eqref{eqn:Pauli}
413: \cite{Sharpe:1998xm,Rupak:2002sm,Bar:2002nr}:
414: \begin{equation}\label{eqn:Lchi}
415:   {\cal L} =
416:    \frac{f^2}{8}
417:     \str\left(\partial_\mu\Sigma^\dagger \partial_\mu\Sigma\right)
418:     - \l_m\,\str\left(m_Q\Sigma^\dagger+m_Q^\dagger\Sigma\right)
419:     - a \l_a\,\str\left(w_Q\Sigma^\dagger+w_Q^\dagger\Sigma\right) 
420: \end{equation}
421: where
422: \begin{equation} \label{eqn:Sigma}
423:   \Sigma=\exp\left(\frac{2i\Phi}{f}\right)
424:   = \xi^2
425: ,\end{equation}
426: \begin{equation}
427:   \Phi=
428:     \left(
429:       \begin{array}{cc}
430:         M & \chi^{\dagger} \\ 
431:         \chi & \tilde{M}
432:       \end{array}
433:     \right)
434: ,\end{equation}
435: $f=132$~MeV, and the str() denotes a graded flavor trace.  
436: The $M$, $\tilde{M}$, and $\chi$ are matrices
437: of pseudo-Goldstone bosons 
438: and pseudo-Goldstone fermions,
439: see, for example,~\cite{Chen:2001yi}.
440: Expanding the Lagrangian in \eqref{eqn:Lchi} one finds that
441: to lowest order mesons with quark content $Q\bar{Q'}$
442: have mass\footnote{%
443: The quark masses $m_Q$ above are not those customarily used on the lattice~\cite{Sharpe:1998xm,Aoki:2003yv}, 
444: because one usually defines the quark mass in terms of a critical parameter for which the meson masses vanish.  
445: If this is indeed the way one defines the renormalized quark mass, 
446: then the parameter $\l_a$ in the Lagrangian can be set to zero and the issue of $a$-dependent 
447: loop-meson masses in baryonic observables will never confront us.  Allowing for other definitions 
448: of the lattice renormalized quark mass to avoid this fine-tuning, we keep $\l_a \neq 0$ and deal with the possibility that 
449: the loop-meson masses remain $a$-dependent. Notice this issue only arises for Wilson quarks.
450: }
451: \begin{equation}\label{eqn:mqq}
452:   m_{QQ'}^2=\frac{4}{f^2} \left[ \l_m (m_Q+m_{Q'}) + a \l_a (w_Q + w_{Q'}) \right]
453: .\end{equation}
454: The flavor singlet field is rendered heavy by the $U(1)_A$ anomaly
455: and has been integrated out in \PQCPT, however,
456: the propagator of the flavor-neutral field deviates from a simple pole 
457: form~\cite{Sharpe:2001fh}. 
458: For $a,b = u,d,j,l,\tilde u,\tilde d$, the leading-order $\eta_a \eta_b$ propagator is given by
459: \begin{equation}
460: {\cal G}_{\eta_a \eta_b} =
461:         \frac{i \epsilon_a \delta_{ab}}{q^2 - m^2_{\eta_a} +i\epsilon}
462:         - \frac{i}{2} \frac{\epsilon_a \epsilon_b \left(q^2 - m^2_{jj}
463:             \right) }
464:             {\left(q^2 - m^2_{\eta_a} +i\epsilon \right)
465:              \left(q^2 - m^2_{\eta_b} +i\epsilon \right)}\, ,
466: \end{equation}
467: where
468: \begin{equation}
469: \epsilon_a = (-1)^{1+\eta_a}
470: .\end{equation}
471: The flavor neutral propagator can be conveniently rewritten as
472: \begin{equation}
473: {\cal G}_{\eta_a \eta_b} =
474:          \e_a \d_{ab} P_a +
475:          \e_a \e_b {\cal H}_{ab}\left(P_a,P_b\right),
476: \end{equation}
477: where
478: \begin{eqnarray}
479:      P_a &=& \frac{i}{q^2 - m^2_{\eta_a} +i\e},\ 
480:      P_b = \frac{i}{q^2 - m^2_{\eta_b} +i\e},\, 
481: \nonumber\\
482: \nonumber\\
483:      {\cal H}_{ab}\left(A,B\right) &=& 
484:            -\frac{1}{2}\left[
485:              \frac{m^2_{\eta_a} - m^2_{jj}}{m^2_{\eta_a} - m^2_{\eta_b}}
486:                  A
487:             -\frac{m^2_{\eta_b} - m^2_{jj}}{m^2_{\eta_b} - m^2_{\eta_a}}
488:                  B \ \right].
489: \label{eq:Hfunction2}
490: \end{eqnarray}
491: 
492: 
493: 
494: 
495: 
496: At $\cO(\e^4)$, one has contributions to the meson Lagrangian from
497: operators of $\cO(p^4)$, $\cO(p^2 \, m_Q)$, and $\cO(m_Q^2)$. These are
498: the Gasser-Leutwyler terms. Additionally there are terms of $\cO( p^2 \, a)$, 
499: $\cO(m_Q \, a)$, and $\cO(a^2)$ that are generalizations of the Gasser-Leutwyler
500: terms for the Symanzik lattice action.  These have been determined 
501: in~\cite{Bar:2003mh,Aoki:2003yv} and we do not duplicate them here. Since our concern 
502: does not lie in the meson sector, mesons will only enter via loop calculations. Retaining
503: the meson masses to $\cO(\e^4)$ in a typical baryon calculation leads to corrections of 
504: $\cO(\e^5)$ or higher. These are beyond the order we work, thus the 
505: Lagrangian in Eq.~\eqref{eqn:Lchi} is sufficient for our purposes.
506: 
507: 
508: 
509: 
510: 
511: 
512: 
513: \subsection{Baryons} \label{s:baryons}
514: 
515: 
516: 
517: Having reviewed the Symanzik Lagrangian through $\cO(a^2)$ and the relevant pieces of meson \PQCPT\ at finite $a$, 
518: we now extend \PQCPT\ in the baryon sector to $\cO(a^2)$. First let us detail the situation at $\cO(a)$. 
519: In $SU(4|2)$ \PQCPT, the spin-$\frac{1}{2}$ baryons 
520: are embedded in the $\bf{70}$-dimensional super-multiplet $\cB^{ijk}$, that
521: contains the nucleons, while the spin-$\frac{3}{2}$ baryons
522: are embedded in the $\bf{44}$-dimensional super-multiplet $\cT_\mu^{ijk}$, that 
523: contains the deltas~\cite{Labrenz:1996jy,Beane:2002vq}. 
524: To $\cO(\e^2)$, the free Lagrangian for the $\cB^{ijk}$ and $\cT^{ijk}_\mu$ fields is given 
525: by~\cite{Labrenz:1996jy,Beane:2002vq,Beane:2003xv}
526: \begin{eqnarray} \label{eqn:L}
527:   {\mathcal L}
528:   &=&
529:    i\left(\ol\cB v\cdot{\mathcal D}\cB\right)
530:   -2\a_M\left(\ol\cB \cB{\mathcal M}_+\right)
531:   -2\b_M\left(\ol\cB {\mathcal M}_+\cB\right)
532:   -2\sigma_M\left(\ol\cB\cB\right)\str\left({\mathcal M}_+\right)
533:                               \nonumber \\
534:   &&
535:   - 2 \a_W\left(\ol\cB \cB{\mathcal W}_+\right)
536:   - 2 \b_W\left(\ol\cB {\mathcal W}_+\cB\right)
537:   - 2 \sigma_W\left(\ol\cB\cB\right)\str\left({\mathcal W}_+\right)
538:                               \nonumber \\
539:   &&+i\left(\ol\cT_\mu v\cdot{\mathcal D}\cT_\mu\right)
540:     +\D\left(\ol\cT_\mu\cT_\mu\right)
541:     +2\g_M\left(\ol\cT_\mu {\mathcal M}_+\cT_\mu\right)
542:     -2\ol\sigma_M\left(\ol\cT_\mu\cT_\mu\right)\str\left({\mathcal M}_+\right) 
543: 				\nonumber \\
544:   &&
545:   + 2 \g_W\left(\ol\cT_\mu {\mathcal W}_+\cT_\mu\right)
546:   - 2 \ol\sigma_W\left(\ol\cT_\mu\cT_\mu\right)\str\left({\mathcal W}_+\right)
547: ,\end{eqnarray}
548: where the mass operator is defined by
549: \begin{equation}
550: {\mathcal M}_\pm = \frac{1}{2}\left(\xi^\dagger m_Q \xi^\dagger\pm\xi m_Q \xi\right)
551: ,\end{equation}
552: and the Wilson operator is defined by\footnote{%
553: Technically the SW coefficient $c_{SW}$, with its possible weak logarithmic dependence on $a$, 
554: also enters into the spurion construction. We are ignoring this dependence and do not distinguish between 
555: spurions of the same form which come from different operators in the Symanzik action largely because at $\cO(a^2)$ 
556: the analysis becomes extremely cumbersome.
557: }
558: \begin{equation} \label{eq:Wilson}
559: \mathcal{W}_\pm = \frac{a \L_{QCD}^2}{2} \left(\xi^\dagger w_Q \xi^\dagger\pm\xi w_Q \xi\right)
560: .\end{equation}
561: Here $\D \sim \e$ is the mass splitting between the $\bf{70}$ and $\bf{44}$ in the chiral limit.
562: The parenthesis notation used in Eq.~\eqref{eqn:L} is that of~\cite{Labrenz:1996jy} and is defined 
563: so that the contractions of flavor indices maintain proper transformations under chiral rotations.
564: Notice that the presence of the chiral symmetry breaking SW operator in Eq.~\eqref{eqn:LPQQCD} 
565: has lead to new $\cO(a)$ operators and new dimensionless constants $\a_W$, $\b_W$, $\sigma_W$, $\g_W$, 
566: and $\ol\sigma_W$ in Eq.~\eqref{eqn:L}. 
567: %Notice in our power counting $a \L_\chi^2 \sim m_Q$. 
568: The Lagrangian describing the interactions of the $\cB^{ijk}$ 
569: and $\cT_\mu^{ijk}$ with the pseudo-Goldstone mesons is
570: \begin{equation} \label{eqn:Linteract}
571:   {\cal L} =   
572: 	  2 \a \left(\ol \cB S_\mu \cB A_\mu \right)
573: 	+ 2 \b \left(\ol \cB S_\mu A_\mu \cB \right)
574: 	- 2{\mathcal H}\left(\ol{\cT}_\nu S_\mu A_\mu \cT_\nu\right) 
575:     	+ \sqrt{\frac{3}{2}}\cC
576:   		\left[
577:     			\left(\ol{\cT}_\nu A_\nu \cB\right)+ \left(\ol \cB A_\nu \cT_\nu\right)
578:   		\right]  
579: .\end{equation}
580: The axial-vector and vector meson fields $A_\mu$ and $V_\mu$
581: are defined by: $ A_\mu=\frac{i}{2}
582: \left(\xi\partial_\mu\xi^\dagger-\xi^\dagger\partial_\mu\xi\right)$  
583: and $V_\mu=\frac{1}{2} \left(\xi\partial_\mu\xi^\dagger+\xi^\dagger\partial_\mu\xi\right)$.
584: The latter appears in  Eq.~\eqref{eqn:L} for the
585: covariant derivatives of $\cB_{ijk}$ and $\cT_{ijk}$ 
586: that both have the form
587: \begin{equation}
588:   ({\mathcal D}_\mu \cB)_{ijk}
589:   =
590:   \partial_\mu \cB_{ijk}
591:   +(V_\mu)_{il}\cB_{ljk}
592:   +(-)^{\eta_i(\eta_j+\eta_m)}(V_\mu)_{jm}\cB_{imk}
593:   +(-)^{(\eta_i+\eta_j)(\eta_k+\eta_n)}(V_\mu)_{kn}\cB_{ijn}
594: .\end{equation}
595: The vector $S_\mu$ is the covariant spin operator~\cite{Jenkins:1991jv,Jenkins:1991es,Jenkins:1991ne}.
596: The interaction Lagrangian in Eq.~\eqref{eqn:Linteract} also receives finite $a$ corrections. 
597: In calculating the octet and decuplet masses, however, these lead to effects that are of $\cO(\e^5)$ or higher. 
598: 
599: 
600: At $\cO(\e^4)$, there are contributions to the \PQCPT\ Lagrangian from two insertions of the mass operator $\cM_+$,
601: and contributions from two insertions of the axial current $A_\mu$. 
602: The former contribute to the baryon masses at tree level, the latter at one-loop level. 
603: These operators have been written down in \cite{Walker-Loud:2004hf,Tiburzi:2004rh,Tiburzi:2005na}.
604: There are also operators with an insertion of $v \cdot A \otimes \cM_-$. These do not contribute to the masses at $\cO(\e^4)$. 
605: To extend baryon \PQCPT\ for mixed lattice actions to $\cO(\e^4)$, we must first include all higher-order operators that are 
606: linear in $a$. Secondly we must map the operators in $\cL^{(6)}$ onto the baryon sector. To achieve the former, 
607: we realize that the relevant operators can be formed from the existing $\cO(a)$ operators by insertion of the mass operator $\cM_+$ 
608: (insertion of a derivative is ruled out because the only possibility is $v \cdot D$ which can be eliminated using
609: the equations of motion~\cite{Arzt:1993gz}). For spin-$\frac{1}{2}$ baryons, there are eleven such operators
610: \begin{eqnarray}
611: \cL &=&
612: -\frac{1}{\L_\chi} 
613: \Bigg[ 
614:  b_1^{WM} (-)^{(\eta_i + \eta_j)(\eta_k + \eta_{k'})} \ol \cB {}^{kji} \{\cM_+, \cW_+ \}^{kk'} \cB^{ijk'} 
615: + 
616: b_2^{WM} \ol \cB {}^{kji} \{\cM_+, \cW_+ \}^{ii'} \cB^{i'jk}  
617: \notag \\
618: && + 
619: b_3^{WM} (-)^{\eta_{i'}(\eta_j + \eta_{j'})} \ol \cB {}^{kji} \cM_+^{ii'} \cW_+^{jj'} \cB^{i'j'k} 
620: + 
621: b_4^{WM} (-)^{\eta_{i}(\eta_j + \eta_{j'})} \ol \cB {}^{kji} \cM_+^{jj'} \cW_+^{ii'} \cB^{i'j'k} 
622: \notag \\
623: && + 
624: b_5^{WM} (-)^{\eta_j \eta_{j'} + 1} \ol \cB {}^{kji} 
625: \left(\cM_+^{ij'} \cW_+^{ji'} + \cW_+^{ij'} \cM_+^{ji'} \right) \cB^{i'j'k} 
626: \notag \\
627: && +
628: b_6^{WM} \left( \ol \cB \cB \cM_+\right) \str (\cW_+) 
629: + 
630: b_7^{WM} \left( \ol \cB  \cM_+ \cB \right) \str (\cW_+)
631: + 
632: b_8^{WM} \left( \ol \cB \cB \right) \str ( \cW_+ \cM_+ ) 
633: \notag \\
634: && + 
635: b_9^{WM} \left( \ol \cB \cB \right) \str( \cW_+ ) \str (\cM_+) 
636: + 
637: b_{10}^{WM} \left( \ol \cB \cB \cW_+ \right) \str (\cM_+)
638: + 
639: b_{11}^{WM} \left( \ol \cB \cW_+ \cB \right) \str (\cM_+)
640: \Bigg],
641: \notag \\
642: \label{eq:BAMs}
643: \end{eqnarray}
644: involving $\cW_+ \otimes \cM_+$. 
645: Additionally there are eleven analogous operators involving the operator combination $\cW_- \otimes \cM_-$, however
646: these terms do not contribute to the baryon masses at this order. 
647: Similarly the eight operators containing $v \cdot A \otimes \cW_-$
648: do not contribute to baryon masses to $\cO(\e^4)$. 
649: 
650: 
651: For the spin-$\frac{3}{2}$ baryons, there are six operators at $\cO( m_Q \, a )$ that contribute to the baryon masses. 
652: These terms are  
653: \begin{eqnarray}
654: \cL &=&
655:  \frac{1}{\L_\chi} \Bigg[ 
656: t_1^{WM} \ol \cT {}^{kji}_\mu \{\cM_+, \cW_+ \}^{ii'} \cT^{i'jk}_\mu  
657: +
658: t_2^{WM} (-)^{\eta_{i'}(\eta_j + \eta_{j'})} \ol \cT  {}^{kji}_\mu \cM_+^{ii'} \cW_+^{jj'} \cT^{i'j'k}_\mu
659: \notag \\ 
660: && + 
661: t_3^{WM} \left( \ol \cT_\mu  \cM_+ \cT_\mu \right) \str (\cW_+)
662: + 
663: t_4^{WM} \left( \ol \cT_\mu \cT_\mu \right) \str ( \cW_+ \cM_+ ) 
664: \notag \\
665: && + 
666: t_5^{WM} \left( \ol \cT_\mu \cT_\mu \right) \str( \cW_+ ) \str (\cM_+) 
667: + 
668: t_{6}^{WM} \left( \ol \cT_\mu \cW_+ \cT_\mu \right) \str (\cM_+)
669: \Bigg].
670: \label{eq:TAMs}
671: \end{eqnarray}
672: Additionally there are six analogous terms involving the operator combination $\cW_- \otimes \cM_-$ and four terms containing
673: $v \cdot A \otimes \cW_-$,  but these operators do not 
674: contribute to the baryon masses at this order. 
675: 
676: 
677: Next we must assess the contribution from the different classes of operators in $\cL^{(6)}$. 
678: Class one operators are quark bilinears that do not break chiral symmetry in the valence and sea sectors. 
679: These operators, however, break the full graded chiral symmetry down to the chiral symmetry group $G$
680: of the mixed lattice action. To describe such operators in the effective theory, 
681: we rewrite the class one operators of the Symanzik Lagrangian in a different form. Instead of describing the
682: bilinears in terms of $w_Q$ and $\ol w_Q$ matrices, we use the linear combinations $w_Q + \ol w_Q$ and 
683: $w_Q - \ol w_Q$. The former combination is just the identity and terms in the effective theory
684: that stem from it are trivial to construct because they are chirally invariant. Those with $w_Q - \ol w_Q$
685: have the flavor symmetry of the mixed action. Let us write out one such term from $\cL^{(6)}$, 
686: \begin{equation}
687: \ol Q \Dslash {}^3 (w_Q - \ol w_Q) Q
688: =
689: \ol Q_L \Dslash {}^3 (w_Q - \ol w_Q) Q_L
690: + 
691: \ol Q_R \Dslash {}^3 (w_Q - \ol w_Q) Q_R 
692: \end{equation}
693: To make these terms $SU(4|2)$ chirally invariant, we introduce two spurions $\cO_1$ and $\cO_2$ transforming as
694: \begin{eqnarray}
695: \cO_1 &\to& L \cO_1 L^\dagger \notag \\
696: \cO_2 &\to& R \cO_2 R^\dagger,  
697: \end{eqnarray}
698: that will be assigned the values $\cO_1 = \cO_2 = w_Q - \ol w_Q$. 
699: In the effective field theory, the operators $\cO = \xi^\dagger \cO_{1} \xi$, and $\xi \cO_2 \xi^\dagger$ both  
700: transform as $U \cO U^\dagger$. Instead of using these operators, we work with 
701: linear combinations of definite parity, namely
702: \begin{equation}
703: \cO_\pm = \frac{1}{2} 
704: \left[ 
705: \xi^\dagger (w_Q - \ol w_Q ) \xi \pm \xi (w_Q - \ol w_Q) \xi^\dagger
706: \right]
707: .\end{equation}
708: Thus class one operators get mapped into the effective theory as terms contained in the Lagrangian
709: \begin{eqnarray} \label{eq:A2s}
710: \cL &=& - a^2 \L_{QCD}^3 \Bigg[ 
711: b_0  \left( \ol \cB \cB \right) 
712: +
713: b_1^\cO  \left( \ol \cB \cB \cO_+ \right)
714: + 
715: b_2^\cO \left( \ol \cB \cO_+ \cB \right)
716: + 
717: b_3^\cO \left( \ol \cB \cB \right) \str \left( \cO_+ \right) \notag \\
718: &&
719: -
720: t_0 \left( \ol \cT_\mu \cT_\mu \right) 
721: - 
722: t_1^\cO \left( \ol \cT_\mu \cO_+ \cT_\mu \right) 
723: - 
724: t_2^\cO \left( \ol \cT_\mu \cT_\mu \right) \str \left( \cO_+ \right)
725: \Bigg].
726: \end{eqnarray}
727: When the action is unmixed, the operator $\cO_+$ is proportional to the identity matrix and hence
728: there is only one independent operator for the $\cB$-field and one for the $\cT^\mu$-field 
729: in Eq.~\eqref{eq:A2s}. 
730: 
731: 
732: 
733: 
734: All class two operators have an insertion of the quark mass matrix $m_Q$. Thus these operators are at least 
735: of $\cO( m_Q \, a^2) = \cO(\e^6)$ in our power counting and can be neglected to the order we are working. 
736: Class three operators are four-quark operators that do not break chiral symmetry in the valence and sea sectors.
737: They do break the $SU(4|2)$ chiral symmetry down to the chiral symmetry $G$ of the mixed action.
738: The requisite spurions for these class three operators are $\cO_1 \otimes \cO_1$, $\cO_1 \otimes \cO_2$, $\cO_2 \otimes \cO_1$, 
739: and $\cO_2 \otimes \cO_2$. Hence operators in the effective theory will involve the products $\cO_+ \otimes \cO_+$ and
740: $\cO_- \otimes \cO_-$. The latter do not contribute to baryon masses to the order we work, while the former
741: baryon operators are contained in the Lagrangian
742: \begin{eqnarray}
743: \cL &=&
744: - a^2 \L_{QCD}^3 \Bigg[
745: b_1^{\cO\cO} (-)^{(\eta_i + \eta_j)(\eta_k + \eta_{k'})} \ol \cB {}^{kji} \left( \cO_+ \cO_+ \right)^{kk'} \cB^{ijk'}
746: +
747: b_2^{\cO\cO} \ol \cB {}^{kji} \left( \cO_+ \cO_+ \right)^{ii'} \cB^{i'jk}
748: \notag \\
749: && + 
750: b_3^{\cO\cO} (-)^{\eta_{i'} ( \eta_j + \eta_{j'})} \ol \cB {}^{kji} \cO_+^{ii'} \cO_+^{jj'} \cB^{i'j'k}
751: + 
752: b_4^{\cO\cO} (-)^{\eta_j \eta_{j'} +1} \ol \cB {}^{kji} \cO_+^{ij'} \cO_+^{ji'} \cB^{i'j'k}
753: \notag \\
754: && + 
755: b_5^{\cO\cO} \left( \ol \cB \cB  \cO_+ \right) \str (\cO_+)
756: + 
757: b_6^{\cO\cO} \left( \ol \cB \cO_+ \cB \right) \str (\cO_+) 
758: +
759: b_7^{\cO\cO} \left( \ol \cB \cB \right) \str (\cO_+ \cO_+) 
760: \notag \\
761: && + 
762: b_8^{\cO\cO} \left( \ol \cB \cB \right) \str (\cO_+) \str (\cO_+) 
763: \Bigg] 
764: ,\label{eq:OOBs}\end{eqnarray}
765: for the spin-$\frac{1}{2}$ fields, and 
766: \begin{eqnarray}
767: \cL &=&
768: a^2 \L_{QCD}^3 \Bigg[
769: t_1^{\cO\cO} \ol \cT {}^{kji}_\mu \left( \cO_+ \cO_+ \right)^{ii'} \cT^{i'jk}_\mu
770: +
771: t_2^{\cO\cO} (-)^{\eta_{i'} ( \eta_j + \eta_{j'})} \ol \cT {}^{kji}_\mu \cO_+^{ii'} \cO_+^{jj'} \cT^{i'j'k}_\mu
772: \notag \\
773: && + 
774: t_3^{\cO\cO} \left( \ol \cT_\mu \cO_+ \cT_\mu \right) \str (\cO_+) 
775: + 
776: t_4^{\cO\cO} \left( \ol \cT_\mu \cT_\mu \right) \str (\cO_+ \cO_+)
777: + 
778: t_5^{\cO\cO} \left( \ol \cT_\mu \cT_\mu \right) \str (\cO_+ ) \str (\cO_+)
779: \Bigg],
780: \notag \\
781: \label{eq:OOTs}
782: \end{eqnarray}
783: for the spin-$\frac{3}{2}$ fields. 
784: When considering unmixed actions, all of the operators in Eqs.~\eqref{eq:OOBs} and \eqref{eq:OOTs} become 
785: redundant compared to those in Eq.~\eqref{eq:A2s}, because $\cO_+$ is proportional to the identity. 
786: 
787: 
788: 
789: 
790: Class four operators are four-quark operators that break chiral symmetry in the Wilson sector of the theory.
791: Such terms involve the flavor structure $w_Q \otimes w_Q$, which must be promoted to various spurions.  
792: In the Symanzik Lagrangian, for example, we have the operator
793: \begin{eqnarray}
794: \left( \ol Q w_Q Q \right)^2 
795: &=&  
796: \left( \ol Q_L w_Q Q_R \right)^2 
797: +  
798: \left( \ol Q_L w_Q Q_R \right) \left( \ol Q_R w_Q Q_L \right) \notag \\
799: && + 
800: \left( \ol Q_R w_Q Q_L \right) \left( \ol Q_L w_Q Q_R \right)
801: +
802: \left( \ol Q_R w_Q Q_L \right)^2
803: ,\end{eqnarray}
804: and so we require the spurions~\cite{Bar:2003mh}
805: \begin{eqnarray}
806: B_1 \otimes B_2 &\to& L B_1 R^\dagger \otimes L B_2 R^\dagger \notag \\
807: B_1^\dagger \otimes B_2^\dagger &\to& R B_1^\dagger L^\dagger \otimes R B_2^\dagger L^\dagger \notag \\
808: C_1 \otimes C_2 &\to& R C_1 L^\dagger \otimes L C_2 R^\dagger \notag \\
809: C_1^\dagger \otimes C_2^\dagger &\to& L C_1^\dagger R^\dagger \otimes R C_2^\dagger L^\dagger, 
810: \end{eqnarray}
811: that will ultimately be given the values $B_1 = B_2 = C_1 = C_2 = a \L_{QCD}^2 w_Q$, and similarly for their
812: Hermitian conjugates. Now the operators $\cO = \xi^\dagger B_{1,2} \xi^\dagger$, $\xi B_{1,2}^\dagger \xi$, 
813: $\xi C_1 \xi$, $\xi^\dagger C_1^\dagger \xi^\dagger$, $\xi^\dagger C_2 \xi^\dagger$, and
814: $\xi C_2^\dagger \xi$ all transform as $U \cO U^\dagger$  under their respective spurion transformations. 
815: Moreover when assigned constant values for their spurions, the operators involving $C$'s and $B_2$'s become indistinguishable from those 
816: involving $B_1$ and $B_1^\dagger$. 
817: Finally instead of working with the operators $\xi w_Q \xi$ and $\xi^\dagger w_Q \xi^\dagger$, we work with linear 
818: combinations that are parity even and odd, which are the $\cW_\pm$ operators, respectively, that were introduced previously.  
819: Thus our spurion analysis shows  
820: terms in the effective theory will involve products of Wilson operators 
821: $\cW_+ \otimes \cW_+$, and  $\cW_- \otimes \cW_-$.\footnote{%
822: As spurions, these have the same effect as squares of the $\cO(a)$ spurions. Thus
823: we need not consider such higher-order effects from lower order terms.
824: } 
825: The latter products do not contribute to the baryon masses at this order. 
826: Thus for the spin-$\frac{1}{2}$ baryons, we have the eight terms from class 
827: four operators
828: \begin{eqnarray}
829: \cL &=&
830: - \frac{1}{\L_{QCD}} \Bigg[
831: b_1^W (-)^{(\eta_i + \eta_j)(\eta_k + \eta_{k'})} \ol \cB {}^{kji} \left( \cW_+ \cW_+ \right)^{kk'} \cB^{ijk'}
832: +
833: b_2^W \ol \cB {}^{kji} \left( \cW_+ \cW_+ \right)^{ii'} \cB^{i'jk}
834: \notag \\
835: && + 
836: b_3^W (-)^{\eta_{i'} ( \eta_j + \eta_{j'})} \ol \cB {}^{kji} \cW_+^{ii'} \cW_+^{jj'} \cB^{i'j'k}
837: + 
838: b_4^W (-)^{\eta_j \eta_{j'} +1} \ol \cB {}^{kji} \cW_+^{ij'} \cW_+^{ji'} \cB^{i'j'k}
839: \notag \\
840: && + 
841: b_5^W \left( \ol \cB \cB  \cW_+ \right) \str (\cW_+)
842: + 
843: b_6^W \left( \ol \cB \cW_+ \cB \right) \str (\cW_+) 
844: +
845: b_7^W \left( \ol \cB \cB \right) \str (\cW_+ \cW_+) 
846: \notag \\
847: && + 
848: b_8^W \left( \ol \cB \cB \right) \str (\cW_+) \str (\cW_+) 
849: \Bigg] 
850: ;\label{eq:WBs}\end{eqnarray}
851: while in the spin-$\frac{3}{2}$ sector, we have five terms
852: \begin{eqnarray}
853: \cL &=&
854: \frac{1}{\L_{QCD}} \Bigg[
855: t_1^W \ol \cT {}^{kji}_\mu \left( \cW_+ \cW_+ \right)^{ii'} \cT^{i'jk}_\mu
856: +
857: t_2^W (-)^{\eta_{i'} ( \eta_j + \eta_{j'})} \ol \cT {}^{kji}_\mu \cW_+^{ii'} \cW_+^{jj'} \cT^{i'j'k}_\mu
858: \notag \\
859: && + 
860: t_3^W \left( \ol \cT_\mu \cW_+ \cT_\mu \right) \str (\cW_+) 
861: + 
862: t_4^W \left( \ol \cT_\mu \cT_\mu \right) \str (\cW_+ \cW_+)
863: + 
864: t_5^W \left( \ol \cT_\mu \cT_\mu \right) \str (\cW_+ ) \str (\cW_+)
865: \Bigg].
866: \notag \\
867: \label{eq:WTs}
868: \end{eqnarray}
869: 
870: 
871: Class five operators break the $O(4)$ rotational symmetry of Euclidean space. 
872: The lowest-order hypercubic invariants that are parity even are: $v_\mu v_\mu v_\mu v_\mu$, 
873: $v_\mu v_\mu S_\mu S_\mu$,  and $S_\mu S_\mu S_\mu S_\mu$. The latter two are both redundant
874: because, suspending the Einstein summation convention, we have 
875: $S_\mu S_\mu = \frac{1}{4} (\delta_{\mu \mu} + v_\mu v_\mu )$.
876: There is an additional hypercubic invariant for the spin-$\frac{3}{2}$ fields since they 
877: carry a vector index. 
878: Furthermore, these class five operators have the chiral symmetry of the group $G$ of the mixed action, not 
879: the full graded chiral symmetry of $\cL^{(4)}$. 
880: As with class one operators, class five operators in the effective theory 
881: require the insertion of the operator $\cO_+$. Thus the $O(4)$ breaking operators are
882: \begin{eqnarray}
883: \cL 
884: &=& 
885: - a^2 \L_{QCD}^3 \Bigg[
886: b^v_0 \, \left( \ol \cB  v_\mu v_\mu v_\mu v_\mu \cB \right)
887: +
888: b^v_1 \, \left( \ol \cB  v_\mu v_\mu v_\mu v_\mu \cB \cO_+\right)
889: +
890: b_2^v \, \left( \ol \cB  v_\mu v_\mu v_\mu v_\mu \cO_+ \cB \right)
891: \notag \\
892: && + 
893: b_3^v \, \left( \ol \cB  v_\mu v_\mu v_\mu v_\mu \cB \right) \str \left( \cO_+ \right) 
894: -
895: t_0^{v} \, \left(  \ol \cT_\nu v_\mu v_\mu v_\mu v_\mu \cT_\nu \right)
896: -
897: t_1^{v} \, \left(  \ol \cT_\nu v_\mu v_\mu v_\mu v_\mu \cO_+ \cT_\nu \right)
898: \notag \\
899: && 
900: -
901: t_2^v \, \left(  \ol \cT_\nu v_\mu v_\mu v_\mu v_\mu \cT_\nu \right) \str \left( \cO_+ \right)
902: - 
903: t_0^{\ol v} \,  \left( \ol \cT_\mu v_\mu v_\mu \cT_\mu \right)
904: - 
905: t_1^{\ol v} \,  \left( \ol \cT_\mu v_\mu v_\mu \cO_+ \cT_\mu \right)
906: \notag \\
907: && 
908: - 
909: t_2^{\ol v} \,  \left( \ol \cT_\mu v_\mu v_\mu \cT_\mu \right) \str \left( \cO_+ \right)
910: \Bigg].
911: \label{eq:vA2s}
912: \end{eqnarray}
913: 
914: 
915: 
916: 
917: The mixing of classes of operators is only possible through linear superpositions. This 
918: is accounted for in the effective theory by linear superpositions of effective operators 
919: and thus need not be addressed separately. 
920: This completes the (nearly) exhaustive listing of $\cO(\e^4)$
921: \PQCPT\ terms for the mixed lattice action. 
922: Despite the large number of operators (over one hundred), many terms do not contribute
923: to baryon masses at tree level and have already been omitted. 
924: Additionally many of the contributing terms are not linearly 
925: independent at leading order. 
926: In the following section we calculate the masses of the nucleons and deltas to $\cO(a^2)$,
927: and thereby determine the number of new free parameters that are required. 
928: 
929: 
930: 
931: 
932: 
933: 
934: \section{Baryon masses to $\cO(a^2)$} \label{s:mass}
935: 
936: 
937: In this section we calculate the masses of the nucleons and deltas in $SU(4|2)$. 
938: Baryon masses in $SU(2)$ are presented in Appendix~\ref{s:2}.
939: 
940: 
941: 
942: 
943: 
944: \subsection{Nucleon mass} \label{s:nm}
945: 
946: In Euclidean lattice field theory, the baryon self energy $\Sigma(\gamma_\mu p_\mu)$ is 
947: no longer rotationally invariant; it is a hypercubic invariant function of the baryon momentum and 
948: gamma matrices. With this functional dependence in mind, we can now calculate the mass of the nucleon. 
949: %
950: %
951: The nucleon mass in the combined lattice spacing and chiral expansion can be written as
952: \begin{eqnarray}
953:      M_{N} = M_0 \left(\mu \right) -  M_{N}^{(2)}\left(\mu \right)
954:                 - M_{N}^{(3)}\left(\mu \right)
955:                 - M_{N}^{(4)}\left(\mu \right) + \ldots
956: \label{eq:Bmassexp}
957: \end{eqnarray}
958: Here, $M_0 \left(\mu \right)$ is the renormalized nucleon
959: mass in the continuum and chiral limits, 
960: which is independent of $m_Q$.
961: $M_{N}^{(n)}$ is the contribution to the mass of order $\e^{n}$, and $\mu$ is the
962: renormalization scale.
963: 
964: 
965: 
966: At order $\e^2$ in the combined lattice spacing and chiral expansion, we have 
967: contributions at tree level from the $\cO(m_Q)$ and $\cO(a)$ operators in Eq.~\eqref{eqn:L}.
968: These contributions to the nucleon mass read
969: \begin{equation}
970: M_N^{(2)} 
971: = 
972: 2 (\a_M  + \b_M ) m_u   + 4  \sigma_M \, m_j  
973: +
974: 2 a \L_{QCD}^2 
975: \left[
976: ( \a_W + \b_W) w_v 
977: + 2 \sigma_W \, w_s 
978: \right] \label{eq:Bmass2}
979: .\end{equation}
980: From this expression, we see that the $\cO(a)$ mass corrections vanish only if
981: both the valence and sea quarks are GW. This is in contrast to the meson sector
982: where only the valence quarks need be GW to ensure the vanishing of linear $a$ contributions.
983: At $\cO(\e^3)$ there are contributions from loop graphs to $M^{(3)}_N$. 
984: These have been determined for $SU(4|2)$ in~\cite{Beane:2002vq} and we 
985: do not duplicate the expressions here because the only modification necessary
986: is to include the $a$-dependence of the loop meson masses via Eq.~\eqref{eqn:mqq}.\footnote{% 
987: Additionally there are local operators that contribute at this order. The form of these operators, 
988: however, is the same as those in Eq.~\eqref{eqn:L} but multiplied by a factor of $\D / \L_\chi$. 
989: We can thus trivially include the effect of these operators by promoting the 
990: LECs to arbitrary polynomial functions of $\D / \L_\chi$ expanded out to the appropriate order.  
991: We do not spell this out explicitly since the determination of these polynomial coefficients requires
992: the ability to vary $\D$. } 
993: 
994: 
995: 
996: \begin{figure}
997: \epsfig{file=amass.eps}
998: \caption{Loop diagrams contributing to the $a$-dependence of the nucleon mass at $\cO(\e^4)$. 
999: Mesons are denoted by a dashed line, flavor neutrals (hairpins) by a crossed dashed line, 
1000: and a thin solid line denotes an octet baryon. The square denotes an $\cO(a)$ vertex. 
1001: }
1002: \label{F:amass}
1003: \end{figure}
1004: 
1005: 
1006: At $\cO(\e^4)$, we have the usual continuum contributions from tree-level $\cO(m_Q^2)$ operators and from 
1007: loop diagrams. These have been detailed for $SU(4|2)$ in \cite{Tiburzi:2005na} and we do not duplicate these
1008: lengthy expressions here. The only modification necessary at finite lattice spacing is the inclusion 
1009: of the $a$-dependence of the loop meson masses from Eq.~\eqref{eqn:mqq}.  The remaining finite lattice spacing corrections
1010: arise at tree level from the $\cO(m_Q \, a)$ operators in Eq.~\eqref{eq:BAMs} and the $\cO(a^2)$ operators in Eqs.~\eqref{eq:A2s}, 
1011: \eqref{eq:WBs}, and \eqref{eq:vA2s}. 
1012: Additionally there are loop diagrams arising from the operators in Eq.~\eqref{eqn:L}, where
1013: either the Wilson operator in Eq.~\eqref{eq:Wilson} is expanded to second order or is inserted
1014: on an internal baryon line. These diagrams are shown in Fig.~\ref{F:amass}. Lastly there are wavefunction
1015: renormalization corrections that are linear in $a$. 
1016: The contribution to the nucleon mass from these operators and loops appears in $M_N^{(4)}$ as a correction of the form
1017: \begin{eqnarray}
1018: \delta M_N^{(4)} 
1019: &=& 
1020: \frac{a \L_{QCD}^2}{\L_\chi} 
1021: \Big[
1022: A  w_v  m_u
1023: + 
1024: B  w_s m_u
1025: + 
1026: C  w_v  m_j
1027: + 
1028: D  w_s  m_j
1029: \Big]
1030: \notag \\
1031: && - 4 \frac{a \L_{QCD}^2}{\L_\chi^2}
1032: \Bigg\{
1033: (\a_W + \b_W)  
1034: \left[ 
1035: (w_v + w_s)
1036: \cL(m_{ju}, \mu)
1037: + 
1038: w_v
1039: \cL(m_{\eta_u}, m_{\eta_u}, \mu)
1040: \right]
1041: \notag \\
1042: && \phantom{spacer} + 
1043: 2 \sigma_W \, w_s
1044: \left[ 
1045: 2 \cL(m_{jj}, \mu)
1046: + 
1047: \cL(m_{\eta_j}, m_{\eta_j}, \mu)
1048: \right]
1049: +
1050: A^N_{ju} \left[ \cL(m_{ju}, \mu) + \frac{2}{3} m_{ju}^2 \right]
1051: \notag \\
1052: && \phantom{spacer} + 
1053: g_{\D N}^2 [B^N_\pi \cJ(m_\pi, \D, \mu) + B^N_{ju} \cJ(m_{ju}, \D, \mu) ]
1054: \Bigg\}
1055: \notag \\
1056: && +
1057: a^2 \L_{QCD}^3 
1058: \Big[E+ E' w_v +E'' w_s+E''' w_v w_s
1059: \notag \\
1060: && \phantom{spacer}
1061: + 
1062: \ol u(p) v_\mu v_\mu v_\mu v_\mu \, u(p)
1063: ( F + F' w_v + F'' w_s )
1064: \Big] \label{eq:Bmass4}
1065: .\end{eqnarray}
1066: The non-analytic functions appearing in this expression are defined by
1067: \begin{eqnarray}
1068: \cL (m_\phi, \mu) &=& m_\phi^2 \log \frac{m_\phi^2}{\mu^2}, \\
1069: \cJ ( m_\phi, \delta, \mu ) &=&
1070: (m_\phi^2 - \d^2) \log \frac{m_\phi^2}{\mu^2} + 2 \d \sqrt{\d^2 - m_\phi^2} \log \left( \frac{\d - \sqrt{\d^2 - m_\phi^2 + i \epsilon}}
1071: {\d + \sqrt{\d^2 - m_\phi^2 + i \epsilon}} \right), \\
1072: %\cM^2(m_\phi, m_{\phi'}) &=& \cH_{\phi \phi'} \left[ m_\phi^2, m_{\phi'}^2 \right] \\
1073: \cL(m_\phi, m_{\phi'}, \mu) &=& \cH_{\phi \phi'} \left[ \cL(m_\phi,\mu), \cL(m_{\phi'},\mu) \right], \\
1074: \cJ(m_\phi, m_{\phi'},\d,\mu) &=& \cH_{\phi \phi'} \left[ \cJ(m_\phi,\d, \mu), \cJ(m_{\phi'},\d, \mu) \right] 
1075: ,\end{eqnarray}
1076: and one should keep in mind the $a$-dependence of the meson masses. 
1077: The parameters $A$--$F$ are 
1078: replacements for particular combinations of LECs.\footnote{%
1079: For clarity in comparing with the operators written down in Sec.~\ref{s:baryons}, these combinations are:
1080: $A = 2 b_1^{WM} + 2 b_2^{WM} +  b_3^{WM}+  b_4^{WM} + 2 b_5^{WM}$, 
1081: $B = 2 b_6^{WM} + 2 b_7^{WM}$, 
1082: $C = 2 b_{10}^{WM} + 2  b_{11}^{WM}$, 
1083: $D = 2 b_8^{WM} + 2 b_9^{WM}$, 
1084: $E = b_0 - b_1^\cO - b_2^\cO - 2 b_3^\cO + b_1^{\cO \cO} + b_2^{\cO \cO} + b_3^{\cO \cO} + b_4^{\cO \cO} + 2 b_5^{\cO \cO} 
1085: + 2 b_6^{\cO \cO} + 2 b_7^{\cO \cO} + 4 b_8^{\cO \cO}$, 
1086: $E' = b_1^W + b_2^W + b_3^W + b_4^W + 2 b_1^\cO + 2 b_2^\cO - 4 b_5^{\cO \cO} - 4 b_6^{\cO \cO}$, 
1087: $E'' = 2 b_7^W + 4 b_8^W + 4 b_3^\cO - 4 b_5^{\cO \cO} - 4 b_6^{\cO \cO}$, 
1088: $E''' = 2 b_5^W + 2 b_6^W + 8 b_5^{\cO \cO} + 8 b_6^{\cO \cO}$, 
1089: $F = b^v_0 - b_1^v - b_2^v - 2 b_3^v$,
1090: $F' = 2 b_1^v + 2 b_2^v$,
1091: and
1092: $F'' = 4 b_3^v$.  
1093: }
1094: The meson loop coefficients $A^N_{ju}$, $B^N_\pi$, and $B^N_{ju}$ are given by
1095: \begin{eqnarray}
1096: A^N_{ju} &=& -\frac{1}{4} \Bigg\{ 2 w_v \left[ - 2 g_A^2 \a_W + g_A g_1 \b_W - \frac{1}{2} g_1^2 (\a_W + 3 \b_W) \right] 
1097: \notag \\
1098: && \phantom{spac}
1099: + w_s \left[ ( 2 g_A^2 + g_A g_1) \a_W + g_1^2 ( \a_W + 3 \b_W)  \right] \Bigg\}, \notag \\
1100: B^N_\pi &=& w_v ( \a_W + \b_W + \gamma_W ) + 2 w_s ( \sigma_W - \ol \sigma_W ), \notag \\
1101: B^N_{ju} &=& w_v \left( \a_W + \b_W + \frac{2}{3} \gamma_W \right) + w_s \left( 2 \sigma_W - 2 \ol \sigma_W + \frac{1}{3} \gamma_W \right) 
1102: .\end{eqnarray}
1103: 
1104: 
1105: Despite the complicated form of Eq.~\eqref{eq:Bmass4}, there are at most only four free parameters depending 
1106: on the mixed action considered. The reason for the complicated form is that the above equation
1107: shows the interplay between all possible theories with mixed actions.  Precisely which operators 
1108: are present or, on the other hand, can be eliminated is perhaps of academic interest because
1109: at this order many of the individual contributions cannot be resolved from lattice data. 
1110: For example, in a theory with Wilson valence and sea quarks, the contribution $A + B$ cannot be further resolved. 
1111: %
1112: %
1113: %
1114: %
1115: %
1116: %
1117: In Table~\ref{t:summary}, we
1118: summarize the number of free parameters introduced by $\cO(a)$, $\cO(m_Q \,a)$, 
1119: and $\cO(a^2)$ operators in the baryon sector for various partially quenched theories. 
1120: Of course, the simplest situation occurs in a theory with GW valence and sea quarks. Here
1121: we see that there are $\cO(a^2)$ corrections to the nucleon self energy of two forms, one
1122: of which violates $O(4)$ symmetry. 
1123: 
1124: 
1125: Let us investigate the violation of $O(4)$ rotational invariance in detail. For an on-shell nucleon
1126: at rest, the $O(4)$ breaking operator merely contributes an additive $a^2$ shift to the mass. 
1127: This correction with $F$ coefficients becomes indistinguishable from the $a^2$ corrections 
1128: which respect $O(4)$ [these have $E$ coefficients in Eq.~\eqref{eq:Bmass4}]. 
1129: For any particle with non-relativistic three-momentum $\bm{p}$, 
1130: however, we can see the effects of $O(4)$ violation in the dispersion relation.
1131: To write the dispersion relation compactly, let us assume both valence and sea quarks are GW.
1132: The same form of the dispersion relation holds for actions involving Wilson quarks with the addition
1133: of various $\cO(a)$ contributions. First let us ignore the contributions from $O(4)$ breaking operators. 
1134: The dispersion relation has the familiar form
1135: \begin{equation} \label{eq:familiar}
1136: E_{\bm{p}} = M(a) + \frac{\bm{p}^2}{2 M(a)} + \ldots
1137: ,\end{equation}
1138: where $M(a) = M(0) - a^2 \L_{QCD}^2 E$, and $M(0)$ denotes the particle mass in the continuum 
1139: limit (which remains quark mass dependent). 
1140: 
1141: 
1142: 
1143: 
1144: To add the $O(4)$ breaking terms to the dispersion relation, we can use a 
1145: non-relativistic expansion for the particle four-velocity or utilize reparameterization 
1146: invariance to deduce the corrections to the dispersion relation. We relegate this 
1147: discussion to Appendix~\ref{s:RPI}, where we address particles with spin less than two. 
1148: Using the result in the Appendix for the spin-$1/2$ nucleon, 
1149: the relation between the particle energy and momentum is then given by
1150: \begin{eqnarray} \label{eq:dispersion}
1151: E_{\bm{p}} 
1152: &=& 
1153: M(0) 
1154: -
1155: a^2 \L_{QCD}^3 (E + F) 
1156: + 
1157: \frac{\bm{p}^2}{2 M(0)} \left[ 1 + \frac{a^2 \L_{QCD}^3}{M(0)} ( E + 5 F ) \right] 
1158: .\end{eqnarray}
1159: Here we have written the non-relativistic expansion to second order and have retained the leading lattice-spacing 
1160: correction for each term. Notice for a particle at rest, we recover the simple result of an $a^2$ shift. 
1161: The kinetic energy receives corrections at $\cO(a^2)$, these are due to the $a^2$ shift in mass as well 
1162: as the $O(4)$ contribution in Eq.~\eqref{eq:vvvv}. If it were not for the latter contribution, the dispersion
1163: relation would retain the form in Eq.~\eqref{eq:familiar} with $M(a) = M(0) - a^2 \L_{QCD}^2 ( E + F)$. 
1164: Additionally the relativistic correction to the kinetic energy will be modified with a coefficient 
1165: proportional to $a^2$ and a different linear combination of $E$ and $F$. Also the energy 
1166: is only a cubic invariant function of the spatial momenta $p_i$. Because four factors of the momentum are required, 
1167: the leading $O(3)$ breaking term in the energy occurs at $\cO(a^2 M^{-4})$ with a coefficient $- F$.   
1168: 
1169: 
1170: 
1171: 
1172: 
1173: 
1174: \subsection{Delta mass} \label{s:dm}
1175: 
1176: 
1177: 
1178: For the masses of spin-$\frac{3}{2}$ baryons, there is an additional feature on a hypercube due to the fact
1179: that the Rarita-Schwinger fields themselves carry a vector index. 
1180: The one-particle irreducible self energy for a spin-$\frac{3}{2}$ field on a hypercube has the form
1181: \begin{equation}
1182: \ol u_\mu(p) \, \Sigma_{\mu \nu} \, u_\nu (p) 
1183: = 
1184: \ol u_\mu(p) \, 
1185: \left( 
1186: \Sigma + \ol \Sigma_{\mu \mu} 
1187: \right) u_\mu (p) 
1188: ,\end{equation}
1189: where $\Sigma$ and $\ol \Sigma_{\mu\mu}$ are both hypercubic invariant functions. 
1190: With this in mind, we can now calculate $\cO(a)$, $\cO(m_Q \, a)$ and $\cO(a^2)$ 
1191: contributions to the delta mass.
1192: %
1193: %
1194: %
1195: %
1196: %
1197: The mass of the delta in the combined lattice spacing and chiral expansion can be written as
1198: \begin{eqnarray}
1199:      M_{\D} = M_0 \left(\mu \right) + \D(\mu) +  M_{\D}^{(2)}\left(\mu \right)
1200:                 + M_{\D}^{(3)}\left(\mu \right)
1201:                 + M_{\D}^{(4)}\left(\mu \right) + \ldots
1202: \label{eq:Tmassexp}
1203: \end{eqnarray}
1204: Here, $M_0 \left(\mu \right)$ is the renormalized baryon mass 
1205: in the continuum and chiral limits and $\D(\mu)$ is the renormalized mass splitting between the spin-$\frac{3}{2}$ and spin-$\frac{1}{2}$
1206: baryons in the continuum and chiral limits. Both of these quantities are independent of $a$ and $m_Q$.
1207: $M_{\D}^{(n)}$ is the contribution of order $\e^{n}$, and $\mu$ is the
1208: renormalization scale.  
1209: 
1210: 
1211: At order $\e^2$ in the expansion, we have contributions to the delta mass 
1212: at tree level from the $\cO(m_Q)$ and $\cO(a)$ operators in Eq.~\eqref{eqn:L}.
1213: These lead to
1214: \begin{equation}
1215: M_\D^{(2)} 
1216: = 
1217: 2 \gamma_M  \, m_u 
1218: - 
1219: 4 \ol \sigma_M \, m_j
1220: + 
1221: 2 a \L_{QCD}^2
1222: \left[
1223: \gamma_W \, w_v
1224: - 
1225: 2 \ol \sigma_W \, w_s
1226: \right] \label{eq:Tmass2}
1227: .\end{equation}
1228: Again in contrast with the mesons,  we note that both valence and sea quarks must be GW for the $\cO(a)$ delta mass corrections to vanish.
1229: At $\cO(\e^3)$ there are contributions from loop graphs to $M^{(3)}_\D$. 
1230: These have been determined for $SU(4|2)$ in~\cite{Tiburzi:2005na} and we 
1231: do not duplicate the expressions here, because the only modification necessary
1232: is to include the $a$-dependence of the loop meson masses via Eq.~\eqref{eqn:mqq}. 
1233: 
1234: 
1235: 
1236: \begin{figure}
1237: \epsfig{file=decmass.eps}
1238: \caption{Loop diagrams contributing to the $a$-dependence of the delta mass. 
1239: Mesons are denoted by a dashed line, flavor neutrals (hairpins) by a crossed dashed line, 
1240: and a thick solid line denotes an decuplet baryon. The square denotes an $\cO(a)$ vertex. 
1241: }
1242: \label{F:decmass}
1243: \end{figure}
1244: 
1245: 
1246: 
1247: Finally at $\cO(\e^4)$ there are the usual contributions from tree-level $\cO(m_Q^2)$ operators and from 
1248: loop diagrams. For the delta, these have been detailed for $SU(4|2)$ in \cite{Tiburzi:2005na}.
1249: The remaining finite lattice-spacing corrections
1250: arise at tree level from the $\cO(m_Q \, a)$ operators in Eq.~\eqref{eq:TAMs} and the $\cO(a^2)$ operators in Eqs.~\eqref{eq:A2s}, 
1251: \eqref{eq:WTs}, and \eqref{eq:vA2s}. Additionally there are loop contributions stemming from the 
1252: operators in Eq.~\eqref{eqn:L}. These loop diagrams are depicted in Fig.~\ref{F:decmass}. Lastly 
1253: we must include wavefunction renormalization corrections that are linear in $a$.  
1254: The contribution to the delta mass from these operators and loops yields
1255: \begin{eqnarray}
1256: \delta M_\D^{(4)} 
1257: &=& 
1258: \frac{a \L_{QCD}^2}{\L_\chi} \Big[
1259: A  w_v   m_u
1260: + 
1261: B  w_s m_u
1262: + 
1263: C w_v m_j
1264: + 
1265: D w_s m_j
1266: \Big]
1267: \notag \\
1268: && 
1269: - 4 \frac{a \L_{QCD}^2}{\L_\chi^2} 
1270: \Bigg\{
1271: \g_W 
1272: \left[ 
1273: (w_v + w_s) \cL(m_{ju}, \mu)
1274: + 
1275: w_v
1276: \cL(m_{\eta_u}, m_{\eta_u}, \mu)
1277: \right]
1278: \notag \\
1279: && \phantom{spacs} \, - 
1280: 2 \ol \sigma_W \, w_s
1281: \left[ 
1282: 2 \cL(m_{jj}, \mu)
1283: + 
1284: \cL(m_{\eta_j}, m_{\eta_j}, \mu)
1285: \right]
1286: \notag \\
1287: && \phantom{spacs}
1288: + \frac{5}{27} \gamma_W g_{\D \D}^2 (w_v - w_s) \left[ \cL (m_{ju},\mu) + \frac{26}{15} m_{ju^2} \right]
1289: \notag \\
1290: && \phantom{spacs}
1291: - 
1292: \frac{1}{2}  g_{\D N}^2 \left[ B^\D_\pi \cJ(m_\pi, -\D, \mu) + B^\D_{ju} \cJ(m_{ju}, - \D, \mu) \right]
1293: \Bigg\}
1294: \notag \\
1295: && +
1296: a^2 \L_{QCD}^3 
1297: \Big[
1298: E 
1299: + 
1300: E'  w_v 
1301: +
1302: E'' w_s
1303: + 
1304: E''' w_v w_s
1305: \notag \\
1306: && \phantom{spacs}
1307: + 
1308: \ol u(p)_\nu v_\mu v_\mu v_\mu v_\mu \, u_\nu(p)
1309: \Big(
1310: F + F' w_v + F'' w_s
1311: \Big)
1312: \notag \\
1313: && \phantom{spacs}
1314: + 
1315: \ol u_\mu (p) v_\mu v_\mu \, u_\mu(p) 
1316: \Big( 
1317: G + G' w_v + G'' w_s
1318: \Big) 
1319:  \Big] \label{eq:Tmass4}
1320: .\end{eqnarray}
1321: The meson loop coefficients $B^\D_\pi$ and $B^\D_{ju}$ are given by
1322: \begin{eqnarray}
1323: B^\D_\pi &=& w_v ( \a_W + \b_W + \gamma_W ) + 2 w_s ( \sigma_W - \ol \sigma_W ),  \notag \\
1324: B^\D_{ju} &=& - \frac{1}{3} [w_v ( 5 \a_W + 2 \b_W + 6 \gamma_W ) + w_s (\a_W + 4 \b_W + 12 \sigma_W - 12 \ol \sigma_W)] \notag
1325: .\end{eqnarray} 
1326: The complicated nature of this expression results from considering general theories with mixed actions. 
1327: The formula retains book-keeping factors that indicate which operators contribute in particular mixed action theories. 
1328: Notice, for example, that for a theory with Wilson valence and sea quarks, the combination $A + B$ 
1329: cannot be further resolved from lattice data. Here the parameters $A$--$G$ depend on particular linear combinations
1330: of the LECs in the spin-$\frac{3}{2}$ sector.\footnote{%
1331: For clarity in comparing with Sec.~\ref{s:baryons}, these combinations of LECs are:
1332: $A = 2 t_1^{WM} +  t_2^{WM}$, 
1333: $B = 2 t_3^{WM}$, 
1334: $C = 2 t_6^{WM}$, 
1335: $D = 2 t_4^{WM} + 4 t_5^{WM}$, 
1336: $E = t_0 - t_1^\cO - 2 t_2^\cO + t_1^{\cO \cO} + t_2^{\cO \cO} +2 t_3^{\cO \cO} + 2 t_4^{\cO \cO} + 4 t_5^{\cO \cO}$, 
1337: $E' = t_1^W + t_2^W + 2 t_1^\cO - 4 t_3^{\cO \cO}$, 
1338: $E'' = 2 t_4^W + 4 t_5^W+ 4 t_2^\cO - 4 t_3^{\cO \cO}$, 
1339: $E''' =2 t_3^W + 8 t_3^{\cO \cO}$, 
1340: $F = t^{v}_0 - t^{v}_1 - 2 t^{v}_2$,
1341: $F' = 2 t^{v}_1$, 
1342: $F'' = 4 t^{v}_2$.
1343: $G = t^{\ol v}_0 - t^{\ol v}_1 - 2 t^{\ol v}_2$, 
1344: $G' = 2 t^{\ol v}_1$,
1345: and
1346: $G'' = 4 t^{\ol v}_2$,
1347: }
1348: The number of free parameters introduced for Wilson valence and sea quarks from
1349: $\cO(a)$, $\cO(m_Q\, a)$, and $\cO(a^2)$ operators is six.
1350: Table~\ref{t:summary} lists the number of free parameters
1351: introduced from finite lattice spacing effects for delta mass in mixed action theories. 
1352: The simplest expression results from an unmixed action of GW quarks. Corrections of 
1353: course start at $\cO(a^2)$, but are of three distinct forms given the possible hypercubic invariants.
1354: 
1355: 
1356: 
1357: As we know from determining the nucleon mass above, the $O(4)$ violating terms proportional to 
1358: $v_\mu v_\mu v_\mu v_\mu$ in Eq.~\eqref{eq:Tmass4} with coefficients $F$ give rise to corrections
1359: to the dispersion relation. There are additional corrections with coefficients $G$ that are tied 
1360: to the Rarita-Schwinger spinors in the form $\ol u_\mu(p) v_\mu v_\mu \, u_\mu(p)$. 
1361: Because of the constraint $v_\mu \, u_\mu(p) = 0$, these new terms
1362: do not contribute for a particle at rest. To consider a particle moving with non-relativistic three momentum $\bm{p}$, 
1363: we must use explicit forms for the Rarita-Schwinger spinors to obtain the modification to the energy-momentum 
1364: relation, see Appendix~\ref{s:RPI}. For GW valence and sea quarks, we find a similar dispersion relation
1365: given in Eq.~\eqref{eq:dispersion} but with an additional term 
1366: \begin{eqnarray} \label{eq:new}
1367: E^\l_{\bm{p}} &=&
1368: M(0) 
1369: +
1370: a^2 \L_{QCD}^3 (E + F) 
1371: + 
1372: \frac{\bm{p}^2}{2 M(0)} \left[ 1 - \frac{a^2 \L_{QCD}^3}{M(0)} ( E - 3 F ) \right] 
1373: \notag \\ 
1374: && +
1375: \frac{a^2 \L_{QCD}^3 G}{3 M(0)^2} 
1376: \left[
1377: |\l| \bm{p}_\perp^2 - \left( 2 |\l| - 3 \right) \bm{p}_\l^2
1378: \right]
1379: ,\end{eqnarray} 
1380: where $\lambda = \pm 3/2, \pm 1/2$ refer to the delta spin states quantized with respect to an axis,
1381: $\bm{p}_\l$ is the momentum along that axis, 
1382: and $\bm{p}_\perp$ is the momentum transverse to that axis. 
1383: We have retained only the leading term
1384: in the non-relativistic expansion and its $a^2$ correction. Unlike the nucleon, the delta
1385: dispersion relation is more sensitive to $O(3)$ violation which enters at $\cO(a^2 M^{-2})$ as opposed
1386: to $\cO(a^2 M^{-4})$. The splitting of spin states is proportional to $|\l|$ because one has 
1387: invariance under rotations of the quantization axis by $\pi$. For an unpolarized state, however,
1388: the $O(3)$ violation in Eq.~\eqref{eq:new} averages out. 
1389: 
1390: 
1391: 
1392: 
1393: 
1394: 
1395: 
1396: \begin{table}
1397: \caption{Summary of the new free parameters entering particle masses in chiral perturbation theory
1398: for non-zero lattice spacing. Listed for \PQCPT\ and \CPT\ are the number of linearly independent
1399: operators contributing to the masses of the pion ($m_\pi$), nucleon ($M_N$), and delta ($M_\D$). These operators
1400: are classified as $\cO(a)$, $\cO(m_Q \,a)$, or $\cO(a^2)$. For each theory
1401: the quarks are grouped by species and for partially quenched theories are grouped by 
1402: pairs of valence and sea quark species. The species are labeled W for Wilson, and GW for Ginsparg-Wilson.}
1403: %\begin{ruledtabular}
1404: \begin{tabular}{c | c c c | c c c | c c c }
1405: \PQCPT\ & \multicolumn{3}{c|}{$m_\pi$} & \multicolumn{3}{c|}{$M_N$} & \multicolumn{3}{c}{$M_\D$} \\
1406: $(v)$ $(s)$ & $\; \cO(a)$ & $\; \cO(m_Q \, a) \;$ & $\; \cO(a^2) \;$ 
1407: 	    & $\; \cO(a)$ & $\; \cO(m_Q \, a) \;$ & $\; \cO(a^2) \;$  
1408:             & $\; \cO(a)$ & $\; \cO(m_Q \, a) \;$ & $\; \cO(a^2) \;$  \\
1409: \hline
1410: W W   & $1$ & $2$ & $1$ & $1$ & $2$ & $2$ & $1$ & $2$ & $3$ \\
1411: W GW  & $1$ & $1$ & $1$ & $1$ & $2$ & $2$ & $1$ & $2$ & $3$ \\
1412: GW W  & $0$ & $1$ & $0$ & $1$ & $2$ & $2$ & $1$ & $2$ & $3$ \\
1413: GW GW & $0$ & $0$ & $0$ & $0$ & $0$ & $2$ & $0$ & $0$ & $3$ \\
1414: \multicolumn{10}{c}{} \\
1415: \hline
1416: $SU(2)$ \CPT\ & $\; \cO(a)$ & $\; \cO(m_Q \, a) \;$ & $\; \cO(a^2) \;$  
1417: 	      & $\; \cO(a)$ & $\; \cO(m_Q \, a) \;$ & $\; \cO(a^2) \;$  
1418:               & $\; \cO(a)$ & $\; \cO(m_Q \, a) \;$ & $\; \cO(a^2) \;$  \\
1419: W      & $1$ & $1$ & $1$ & $1$ & $1$ & $2$ & $1$ & $1$ & $3$ \\
1420: GW     & $0$ & $0$ & $0$ & $0$ & $0$ & $2$ & $0$ & $0$ & $3$ \\
1421: \end{tabular}
1422: %\end{ruledtabular}
1423: \label{t:summary}
1424: \end{table}
1425: 
1426: 
1427: 
1428: 
1429: 
1430: \section{\label{s:summy}Summary}
1431: 
1432: 
1433: Above we have extended heavy baryon chiral perturbation theory for the Wilson action to $\cO(a^2)$. 
1434: In our power counting, we have included all operators that are at least of $\cO(\e^4)$, 
1435: this includes operators of $\cO(m_Q \, a)$ and $\cO(a^2)$. We have considered partially quenched $SU(4|2)$ 
1436: with a mixed action, as well as $SU(2)$ in Appendix~\ref{s:2}.
1437: To this order in the combined lattice spacing and
1438: chiral expansion, we saw the necessity for the introduction of a large number of new operators. For the case
1439: of a partially quenched theory with a mixed lattice action, there are well over one hundred operators in the
1440: free Lagrangian containing the spin-$\frac{1}{2}$ and spin-$\frac{3}{2}$ baryons. 
1441: 
1442: 
1443: Despite the introduction of a large number of terms with undetermined LECs, observables calculated to $\cO(a^2)$ 
1444: depend on still relatively few independent parameters. This is because many of the operators introduced
1445: for mixed lattice actions act differently depending on the number of sea quarks contained in particular states of the multiplet. 
1446: Each level of the multiplet must be treated differently in the effective theory because there are no symmetry transformations
1447: between the valence and sea sectors of the theory. This distinction is important if the level-dependent 
1448: operators act on baryons within loops and such a situation only occurs at higher orders in the chiral expansion. 
1449: Moreover, most of the remaining operators for baryons which consist of only valence quarks  
1450: are not independent at leading order and hence 
1451: cannot have their LECs disentangled from lattice data. To make these points clear, we have calculated 
1452: the finite lattice-spacing corrections to baryon masses in various theories. In Table~\ref{t:summary},
1453: we summarize the number of free parameters due to lattice spacing artifacts 
1454: that enter into expressions for baryon masses and can be determined from lattice data.  
1455: The table is separated for the pion, nucleon, and delta into contributions from independent $\cO(a)$, 
1456: $\cO(m_Q \, a)$ and $\cO(a^2)$ operators. 
1457: 
1458: 
1459: There are some further points to observe about the baryon sector at $\cO(a^2)$. 
1460: While there are considerably more operators in the partially quenched theories for mixed actions
1461: as compared to the unquenched theories, the number of independent parameters entering 
1462: into the determination of baryon masses from lattice data is the same for $\cO(a)$ and $\cO(a^2)$ operators;
1463: there is only one additional $\cO(m_Q \, a)$ parameter which stems from the different sea quark mass. 
1464: Nonetheless, the computational benefits of partially quenching are not hindered by the explosion 
1465: in the number of new baryon operators because most are not independent to $\cO(\e^4)$. 
1466: 
1467: 
1468: On the other hand, based on studies in the meson sector~\cite{Bar:2003mh}, one would expect
1469: the number of free parameters to be reduced when considering theories with Ginsparg-Wilson
1470: valence quarks and Wilson sea quarks compared to the Wilson action, \emph{cf}. Table~\ref{t:summary}. 
1471: While there is a clear and sizable reduction in the number 
1472: of baryon operators for actions employing Ginsparg-Wilson valence quarks and Wilson sea quarks
1473: compared to the unmixed Wilson action, 
1474: there is no reduction in the number of free parameters involved in the baryon masses. 
1475: The only reduction in the number of free parameters comes about in simulations employing Ginsparg-Wilson 
1476: quarks in both the valence and sea sectors. Another notable difference for heavy baryon fields is that operators which
1477: reduce the $O(4)$ symmetry to the hypercubic group are present in the effective theory. Such operators were suppressed 
1478: in the pseudoscalar meson sector because the absence of a large mass scale implies
1479: the pseudoscalar momenta are small. For heavy baryon fields, however, there is a large mass scale and the momenta
1480: are automatically large, thus such operators are not suppressed when multiplied by powers of the lattice spacing.  
1481: Similar operators must then enter at $\cO(\e^4)$ for other heavy particles. 
1482: We have spelled out the form of $O(4)$ breaking corrections for particles of spin less than two in Appendix~\ref{s:RPI}. 
1483: 
1484: 
1485: In this work we have seen that near the continuum limit, the spin-$\frac{1}{2}$ baryon self energy has the 
1486: behavior\footnote{For GW quarks, the coefficients $\a$, $\b$, and $\gamma$ are identically zero.}
1487: \begin{equation}
1488: M(a) = M(0) + \a \, a + \b \, a \, m_q \, \log m_q + \g \, a \, m_q + \d \, a^2 
1489: + \epsilon \, \ol u(p) \, v_\mu v_\mu v_\mu v_\mu \, u(p) + \ldots
1490: ,\end{equation}
1491: where the omitted terms are of order $\e^5$ and higher. 
1492: In writing the above expression, we have assumed for simplicity that one fine tunes the quark mass so that the pion mass vanishes
1493: in the chiral limit. In this case there are only polynomial corrections in $a$. The influence of the $O(4)$ breaking term  
1494: on the baryon energy was determined in Sec.~\ref{s:nm}.
1495: For spin-$\frac{3}{2}$ baryons, the behavior of the self energy has the same form, however there is an additional 
1496: contribution to the self energy of the form
1497: \begin{equation}
1498: \ol u_\mu(p) \, \ol \Sigma_{\mu \mu} \, u_\mu(p) = \omega \, a^2 \, \ol u_\mu(p) \, v_\mu v_\mu \, u_\mu(p) + \ldots
1499: ,\end{equation}
1500: and we have detailed the form of this correction in Sec.~\ref{s:dm}. 
1501: A similar such additional term is present for vector mesons as well.
1502: 
1503: 
1504: Knowledge of the quark mass dependence of baryon observables is crucial to perform the
1505: chiral extrapolation of lattice data, extract physical LECs, and make predictions for QCD. 
1506: Artifacts of approximating spacetime on a discrete lattice make the chiral extrapolation 
1507: more challenging because of the introduction of additional error. With \CPT\ and \PQCPT\ 
1508: formulated for the Symanzik action, one can parametrize the dependence on the lattice
1509: spacing and considerably reduce the uncertainty surrounding lattice artifacts. 
1510: The formalism set up here is straightforward to extend to other Wilson-type fermion actions,
1511: twisted mass QCD~\cite{Walker-Loud:2005bt}, for example. 
1512: On the other hand, due to the proliferation of operators, it is doubtful that partially 
1513: quenched baryon staggered \CPT\ would provide much analytic insight at $\cO(a^2)$. 
1514: 
1515: 
1516: 
1517: 
1518: 
1519: 
1520: 
1521: \begin{acknowledgments}
1522: We are indebted to Martin Savage for alerting us to \cite{Savage}. 
1523: We would also like to acknowledge Shailesh Chandrasekharan, Will Detmold,  and Andr\'e Walker-Loud for helpful discussions.
1524: This work is supported in part by the U.S.\ Department of Energy under 
1525: Grant No.\ DE-FG02-96ER40945, and we thank the Institute for Nuclear Theory
1526: at the University of Washington for its hospitality during the final stages of this work. 
1527: \end{acknowledgments}
1528: 
1529: 
1530: 
1531: 
1532: 
1533: \appendix 
1534: 
1535: 
1536: 
1537: \section{Baryon masses to $\cO(a^2)$ in $SU(2)$ \CPT} \label{s:2}
1538: 
1539: In this Appendix, we detail the case of finite lattice-spacing corrections 
1540: to baryon masses in $SU(2)$ \CPT. First we write down all finite lattice spacing terms that arise
1541: to $\cO(\e^4)$ from the Symanzik Lagrangian. Next we determine the corrections
1542: from these operators to the masses of the nucleon and delta. 
1543: 
1544: 
1545: 
1546: The form of the Symanzik Lagrangian for $SU(2)$ flavor is mainly the same 
1547: as in the main text. Of course there is no separation of the 
1548: theory into valence and sea sectors.  The mass matrix for two light flavors
1549: in the isospin limit is given by $m_Q = \diag (m_u , m_u)$, while the Wilson matrix appears as
1550: $w_Q = \diag (w_v, w_v)$. The subscript $v$ now has no particular significance 
1551: in $SU(2)$ \CPT\ other than to maintain the definition of $w_v$ used in above. 
1552: 
1553: 
1554: 
1555: At zero lattice spacing and zero quark mass, two flavor QCD has an $SU(2)_L \otimes SU(2)_R \otimes U(1)_V$
1556: chiral symmetry that is broken down to $SU(2)_V \otimes U(1)_V$. \CPT\ is the low-energy 
1557: effective theory that emerges from perturbing about the physical vacuum. 
1558: The pseudo-Goldstone bosons can be described by a Lagrangian that takes into account
1559: the two sources of chiral symmetry breaking and is given in Eq.~\eqref{eqn:Lchi}, 
1560: with the exception that $\Phi$ is an $SU(2)$ matrix containing just the familiar pions, and 
1561: the $\str$'s are now $\tr$'s. Their masses are given in Eq.~\eqref{eqn:mqq}. 
1562: 
1563: 
1564: In $SU(2)$ \CPT, the delta baryons are contained in the flavor tensor $T^{ijk}_\mu$ which 
1565: is embedded in the \PQCPT\ tensor $\cT^{ijk}_\mu$ simply as $\cT^{ijk}_\mu = T^{ijk}_\mu$,  when all indices
1566: are restricted to $1$--$2$. Consequently the form of the \CPT\ Lagrangian for the delta fields 
1567: has a form very similar to that of \PQCPT.
1568: The nucleons, however, are conveniently described by an $SU(2)$ doublet
1569: \begin{equation}
1570: N = 
1571: \begin{pmatrix}
1572: p \\
1573: n
1574: \end{pmatrix}
1575: .\end{equation}
1576: These states are contained in the \PQCPT\ flavor tensor $\cB^{ijk}$
1577: as~\cite{Beane:2002vq}
1578: \begin{equation}
1579: \cB^{ijk} = \frac{1}{\sqrt{6}} 
1580: \left( 
1581: \e^{ij} N^{k} + \e^{ik} N^{j}
1582: \right)
1583: ,\end{equation}
1584: when all of the indices are restricted to $1$--$2$. Consequently operators involving the nucleons
1585: will appear differently in $SU(2)$. 
1586: 
1587: 
1588: 
1589: 
1590: To $\cO(\e^2)$ the free Lagrangian for the nucleons and deltas reads
1591: \begin{eqnarray}
1592: \cL 
1593: &=& 
1594: i \ol N  v \cdot D N 
1595: - 
1596: 2 \sigma
1597: \ol N N \tr (\cM_+)
1598: - 
1599: 2 \sigma_w 
1600: \ol N N \tr ( \cW_+ ) 
1601: \notag \\
1602: && +
1603: i \ol T_\mu v \cdot D T_\mu 
1604: + 
1605: \D \ol T_\mu T_\mu 
1606: - 
1607: 2 \ol \sigma \, \ol T_\mu T_\mu \tr (\cM_+)
1608: -
1609: 2 \ol \sigma_w \ol T_\mu T_\mu \tr (\cW_+)
1610: .\end{eqnarray}
1611: The LECs for the nucleons and deltas in $SU(4|2)$ are related by the matching equations
1612: \begin{eqnarray}
1613: \sigma     &=&   \frac{1}{2} ( \a_M + \b_M) + \sigma_M, \notag \\
1614: \ol \sigma &=& - \frac{1}{2} \gamma_M + \ol \sigma_M 
1615: ,\end{eqnarray}
1616: and completely analogous equations that relate 
1617: $\sigma_w$, and $\ol \sigma_w$ to the \PQCPT\ parameters 
1618: $\a_W$, $\b_W$, $\sigma_W$, $\gamma_W$, and $\ol \sigma_W$. 
1619: The interaction Lagrangian of $SU(2)$ appears as
1620: \begin{equation}
1621: \cL = 2 g_A \ol N S \cdot A N + 2 g_{\D N} \left( \ol T_\mu A_\mu N + \ol N A_\mu T_\mu \right)
1622: - 2 g_{\D \D} \ol T_\mu S \cdot A T_\mu  
1623: .\end{equation}
1624: The relation of the parameters appearing in the \PQCPT\ Lagrangian in the main text to those in \CPT\
1625: can be found from matching. One finds \cite{Beane:2002vq}: $g_A = \frac{2}{3} \a_M - \frac{1}{3} \b_M$, 
1626: $g_{\D N} = - \cC$, and $g_{\D \D} = \cH$.   
1627: 
1628: 
1629: The first lattice spacing corrections at $\cO(\e^4)$ arise from the $\cO( m_Q \, a)$ operators. In degenerate $SU(2)$, we have
1630: only two such baryon operators
1631: \begin{eqnarray}
1632: \cL 
1633: &=&
1634: - \frac{1}{ \L_\chi} \Bigg[
1635: n^{WM} \, \ol N N \, \tr( \cW_+) \, \tr (\cM_+)  
1636: -
1637: t^{WM} \, \ol T_\mu T_\mu \, \tr( \cW_+) \, \tr (\cM_+)
1638: \Bigg]
1639: .\end{eqnarray}
1640: The remaining terms are of order $\cO(a^2)$ and appear in the Lagrangian
1641: \begin{eqnarray}
1642: \cL &=&
1643: - \frac{1}{\L_{QCD}} \Bigg[
1644: n^W \, \ol N N \, \tr(\cW_+) \tr (\cW_+ ) 
1645: -
1646: t^W \, \ol T_\mu T_\mu \, \tr(\cW_+) \tr (\cW_+ ) 
1647: \Bigg] \notag \\
1648: && - 
1649: a^2 \L_{QCD}^3 \Big[ n \, \ol N N + n^v \ol N v_\mu v_\mu v_\mu v_\mu N 
1650: - t \, \ol T_\mu T_\mu - t^v \, \ol T_\nu v_\mu v_\mu v_\mu v_\mu T_\nu 
1651: - t^{\ol v} \, \ol T_\mu v_\mu v_\mu T_\mu \Big]
1652: .\end{eqnarray}
1653: The coefficients of operators in the $SU(2)$ theory are contained in $SU(4|2)$, but 
1654: because $SU(4|2)$ contains considerably more operators, the matching yields 
1655: algebraically cumbersome expressions that relate the LECs.  
1656: 
1657: 
1658: Having written down all of the relevant operators for the nucleon and delta masses, we now calculate their lattice
1659: spacing corrections. We do not write down the loop contributions from $a$-independent operators because the only 
1660: modification necessary is to include the lattice spacing dependence of the meson masses that appears in Eq.~\eqref{eqn:mqq}. 
1661: For the mass of the nucleon, at $\cO(\e^2)$ we find
1662: \begin{equation}
1663: M_N^{(2)} = 4 \sigma \, m_u  
1664: + 4 a \L_{QCD}^2 \sigma_w \, w_v
1665: ,\end{equation}
1666: The finite lattice spacing corrections to $M_N^{(4)}$ read
1667: \begin{eqnarray}
1668: \delta M_N^{(4)} 
1669: &=& 
1670: 4 w_v \frac{a \L_{QCD}^2}{\L_\chi^2} \Big\{
1671: \L_\chi \, n^{WM} m_u
1672: - 
1673: 3 \sigma_w \cL(m_\pi, \mu)
1674: +
1675: 4 g_{\D N}^2 (\ol \sigma_w - \sigma_w ) 
1676: \left[ \cJ(m_\pi, \D, \mu) + m_\pi^2  \right]
1677: \Big\}
1678: \notag \\
1679: && +
1680: a^2 \L_{QCD}^3 
1681: \Big( n + 4 n^W w_v + n^v \ol u(p) v_\mu v_\mu v_\mu v_\mu \, u(p) \Big)
1682: .\end{eqnarray}
1683: %
1684: %%
1685: %
1686: For the deltas, similar calculations yield the $\cO(\e^2)$ result
1687: \begin{equation}
1688: M_\D^{(2)} 
1689: = 
1690: - 
1691: 4 \ol \sigma \, m_u
1692: -
1693: 4 a \L_{QCD}^2 
1694: \ol \sigma_w \, w_v 
1695: \label{eq:Tmass6}
1696: .\end{equation}
1697: At $\cO(\e^4)$ the remaining finite lattice spacing corrections to delta mass are
1698: \begin{eqnarray}
1699: \delta M_\D^{(4)} 
1700: &=& 
1701: 4 w_v \frac{a \L_{QCD}^2}{\L_\chi^2} \Big[ 
1702: \L_\chi  \, t^{WM} m_u
1703: + 
1704: 3 \ol \sigma_w \cL(m_\pi, \mu)
1705: + 
1706: g_{\D N}^2 (\ol \sigma_w -  \sigma_w) \cJ(m_\pi , - \D, \mu)
1707: \Big]
1708: \notag \\
1709: && +
1710: a^2 \L_{QCD}^3 
1711: \Big( 
1712: t + 4 t^{W} \, w_v + t^v \ol u_\nu(p) v_\mu v_\mu v_\mu v_\mu \, u_\nu(p) + t^{\ol v} \ol u_\mu(p) v_\mu v_\mu \, u_\mu(p)
1713: \Big)
1714: .\label{eq:Tmass7}
1715: \end{eqnarray}
1716: 
1717: 
1718: 
1719: 
1720: \section{$O(4)$ breaking operators} \label{s:RPI}
1721: 
1722: 
1723: Above we addressed $\cO(a^2)$ corrections in the baryon sector. At this order
1724: we saw that $O(4)$ breaking operators are present in the chiral effective theory. 
1725: In this Appendix, we detail the form of these corrections in heavy particle effective
1726: theories for particles of spin less than two. 
1727: 
1728: 
1729: We consider the simplest case first, that of a heavy scalar field $\phi$. 
1730: At $\cO(a^2)$, we have an $O(4)$ breaking operator of the form
1731: \begin{equation} \label{eq:scalar}
1732: \cL = a^2 c_0 \, \phi^\dagger \, v_\mu v_\mu v_\mu v_\mu \, \phi
1733: .\end{equation}
1734: We can either treat the four-velocity $v^\mu$ as the physical velocity of the $\phi$ field, or 
1735: consider the four-velocity to be fixed in the rest frame, for example. Clearly when
1736: the particle is at rest, both descriptions are identical; and with $v_\mu = (0,0,0,i)$, 
1737: we obtain a trivial $a^2$ dependent shift to the $\phi$ mass. Away from rest the effects 
1738: of $O(4)$ breaking are no longer trivial. 
1739: 
1740: First let us consider $v_\mu$ to be the physical velocity of the $\phi$. In terms of 
1741: the particle's spatial momentum $\bm{p}$, we have $v_\mu = (\gamma \bm{p}/M, i \gamma)$, 
1742: where $M$ is the mass of the $\phi$ and $\gamma$ is the Lorentz-Fitzgerald contraction factor. 
1743: In a non-relativistic expansion, we can express the particle's four-velocity in terms of the 
1744: spatial momentum. This enables us to write
1745: \begin{equation} \label{eq:vvvv}
1746: \sum_{\mu = 1}^{4} v_\mu v_\mu v_\mu v_\mu  
1747: = 
1748: 1 
1749: + 
1750: 2 \frac{\bm{p}^2}{M^2} 
1751: + 
1752: \frac{1}{M^4} \left( \sum_{i=1}^{3} p_i p_i p_i p_i + \bm{p}^4 \right)
1753: + 
1754: \cO(M^{-6})
1755: .\end{equation}
1756: Thus the operator in Eq.~\eqref{eq:scalar}, besides shifting the mass by $a^2 c_0$, modifies
1757: the particle's kinetic energy as well as the relativistic correction to the kinetic energy. 
1758: The energy of the $\phi$ is now a cubic invariant function of the spatial momentum.
1759: The breaking of $O(3)$ rotational symmetry for a particle's energy-momentum relation
1760: is, however, small for non-relativistic momenta. 
1761: 
1762: 
1763: Extending the above analysis to non-zero spin involves boosting the various spin
1764: wavefunctions. On the other hand, if one treats the four-velocity as fixed and one can 
1765: derive a string of fixed coefficient operators using reparameterization invariance ~\cite{Luke:1992cs}.
1766: For a fixed four-velocity, we write
1767: the heavy scalar momentum $P$ as $P_\mu = M v_\mu + p_\mu$. This decomposition is arbitrary as one can make 
1768: a reparameterization of the form
1769: \begin{equation}
1770: \begin{cases}
1771: v_\mu \to v_\mu +  q_\mu / M \notag \\
1772: k_\mu \to k_\mu - q_\mu
1773: \end{cases}
1774: \label{eq:RPS}
1775: .\end{equation}
1776: Accordingly the field $\phi(x)$ changes to $e^{i q_\mu x_\mu} \phi(x)$, and
1777: to maintain the normalization of the new four-velocity, we require $v \cdot q = - q^2 / (2 M)$.
1778: The physics is unchanged by this transformation and this is referred to as reparameterization invariance. 
1779: In order for the Lagrangian in Eq.~\eqref{eq:scalar} to be reparameterization invariant (RPI), there must 
1780: be additional operators with fixed coefficients. These can all be written compactly in the manifestly
1781: RPI form
1782: \begin{equation} \label{eq:scalarRPI}
1783: \cL = a^2 c_0 \, \phi^\dagger \, 
1784: \left(v_\mu + \frac{i D_\mu}{M}\right)
1785: \left(v_\mu + \frac{i D_\mu}{M}\right)
1786: \left(v_\mu + \frac{i D_\mu}{M}\right)
1787: \left(v_\mu + \frac{i D_\mu}{M}\right)
1788: \phi
1789: .\end{equation}
1790: %where $\partial_{\perp,\mu} = \partial_\mu + v_\mu v \cdot \partial$. 
1791: Expanding out the terms of Eq.~\eqref{eq:scalarRPI} and acting between states of residual momentum $p_\mu$
1792: taken in the non-relativistic limit, we recover the result of Eq.~\eqref{eq:vvvv}. 
1793: 
1794: 
1795: 
1796: For a heavy spin-$1/2$ particle $\psi$, we have the $O(4)$ breaking Lagrangian
1797: \begin{equation} \label{eq:half}
1798: \cL = a^2 c_{1/2} \ol \psi \, v_\mu v_\mu v_\mu v_\mu \, \psi
1799: ,\end{equation}
1800: which is not RPI. Unlike the scalar case, we have a further constraint on the field
1801: due to the spin degrees of freedom. To maintain the condition $- i \rlap \slash v \psi = \psi$, 
1802: the field $\psi(x)$ must change 
1803: %to $e^{i q_\mu x_\mu} [ 1 - i \rlap \slash q / (2 M)] \psi(x)$ 
1804: under the reparameterization in Eq.~\eqref{eq:RPS}.  
1805: To order $M^{-2}$, we have reparameterization invariant spinor $\Psi$ given by~\cite{Manohar:1997qy} 
1806: \begin{equation}
1807: \Psi = \left( 1 - \frac{\Dslash_\perp}{2 M } - \frac{D^2}{8 m^2} \right) \psi
1808: ,\end{equation}
1809: where $D_{\mu,\perp} = D_\mu + v_\mu v_\nu D_\nu$.  
1810: Since the bilinear $\ol \Psi \Psi = \ol \psi \psi + \cO(M^{-3})$, the RPI form of Eq.~\eqref{eq:half} is
1811: \begin{equation} \label{eq:halfRPI}
1812: \cL = a^2 c_{1/2} \ol \psi 
1813: \left(v_\mu + \frac{i D_\mu}{M}\right)
1814: \left(v_\mu + \frac{i D_\mu}{M}\right)
1815: \left(v_\mu + \frac{i D_\mu}{M}\right)
1816: \left(v_\mu + \frac{i D_\mu}{M}\right)
1817: \psi
1818: ,\end{equation}
1819: and hence the correction to the energy to $\cO(M^{-2})$ has the same form as in Eq.~\eqref{eq:vvvv}.
1820: If we were to write the covariant spinor to fourth order, we could ascertain the coefficient of
1821: the $\bm{p}^4/M^4$ term in Eq.~\eqref{eq:vvvv} for a spin-$1/2$ particle. 
1822: The $O(3)$ breaking term, however, is unchanged. 
1823: 
1824: 
1825: 
1826: 
1827: 
1828: 
1829: The correction for a heavy vector particle, the rho meson $\rho_\mu$ for example, is deduced similarly.
1830: The $\cO(a^2)$ Lagrangian contains two terms 
1831: \begin{equation} \label{eq:rho}
1832: \cL = a^2 c_1 \, \rho_\nu^\dagger \, v_\mu v_\mu v_\mu v_\mu \, \rho_\nu 
1833: + 
1834: a^2  \ol c_1 \, \rho^\dagger_\mu \, v_\mu v_\mu \, \rho_\mu 
1835: ,\end{equation}
1836: and hence there is a new type of contribution to the energy proportional to $\epsilon^*_\mu (p) \, v_\mu v_\mu \, \epsilon_\mu (p)$, 
1837: where $\epsilon_\mu(p)$ is the rho's polarization vector. This correction vanishes for a rho at rest. 
1838: The terms in Eq.~\eqref{eq:rho} are not RPI. To deduce the RPI form of the Lagrangian, we must now also 
1839: maintain the constraint $v_\mu \rho_\mu = 0$. We can build the RPI Lagrangian from RPI vector field $R_\mu$.
1840: %
1841: %transform the field $\rho_\mu(x)$ according to $e^{- i q_\mu x_\mu} (\d_{\mu \nu} - v_\mu q_\nu / M ) \rho_\nu(x)$
1842: %in order to preserve $v_\mu \rho_\mu = 0$ under the reparametrization Eq.~\eqref{eq:RPS}.
1843: To $\cO(M^{-2})$, we have 
1844: \begin{equation} \label{eq:rhocov}
1845: R_\mu = \left(\d_{\mu \nu} - v_\mu \frac{i D_\nu }{M} + \frac{D_\mu D_\nu}{M^2}\right) \rho_\nu
1846: ,\end{equation}
1847: and hence the RPI form of the Lagrangian in Eq.~\eqref{eq:rho} can be deduced from combinations of $R^\dagger_\mu \Gamma R_\mu$. 
1848: Because we are addressing the corrections at tree level, the states are on-shell and 
1849: the additional terms in Eq.~\eqref{eq:rhocov} vanish. In loops the off-shell degrees of freedom
1850: propagate and the additional terms are necessary. 
1851: The on-shell RPI Lagrangian is thus given by
1852: \begin{eqnarray} \label{eq:one}
1853: \cL &=& a^2 c_1 \, \rho_\nu^\dagger 
1854: \left(v_\mu + \frac{i D_\mu}{M}\right)
1855: \left(v_\mu + \frac{i D_\mu}{M}\right)
1856: \left(v_\mu + \frac{i D_\mu}{M}\right)
1857: \left(v_\mu + \frac{i D_\mu}{M}\right)
1858: \rho_\nu \notag \\
1859: && + 
1860: a^2 \ol c_1 \, \rho_\mu^\dagger 
1861: \left(v_\mu + \frac{i D_\mu}{M}\right)
1862: \left(v_\mu + \frac{i D_\mu}{M}\right)
1863: \rho_\mu
1864: .\end{eqnarray}
1865: 
1866: 
1867: We can use the above Lagrangian to determine the $\cO(a^2)$ corrections
1868: to the vector particle's dispersion relation. Let us denote $M(a)$ 
1869: as the particle mass including lattice discretization contributions. 
1870: One such contribution is $a^2 c_1$ that arises from the above Lagrangian. There 
1871: are many others, however, that we have not spelled out. Modification of the
1872: dispersion relation only arises from the terms in Eq.~\eqref{eq:one}.
1873: Utilizing the non-relativistic expansion for the momentum $\bm{p}$ 
1874: and explicit forms of the polarization vectors, 
1875: we arrive at the energy of a spin-one particle
1876: \begin{equation}
1877: E^\l_{\bm{p}} = M(a) + \frac{\bm{p}^2}{2 M(a)} \left( 1 + \frac{4 a^2 c_1}{M(0)}\right) +
1878: \frac{a^2 \ol c_1}{2 M(0)^2} \left[ |\l| \bm{p}_\perp^2 + 2 (|\l| - 1) \bm{p}_\l^2 \right]
1879: ,\end{equation}
1880: where $M(0)$ is the rho mass in the continuum limit, $\lambda = \pm 1, 0$ is the 
1881: spin along an axis, $\bm{p}_\l$ is the momentum along that axis, 
1882: and $\bm{p}_\perp$ is the momentum transverse to that axis.
1883: For a vector particle of given polarization, the energy-momentum relation breaks
1884: $O(3)$ rotational invariance at $\cO(a^2 M^{-2})$. Notice for an unpolarized state,
1885: the rotational symmetry breaking from this term is averaged out and then does not enter until $\cO(a^2 M^{-4})$
1886: as with scalar and spin-$1/2$ particles. 
1887: 
1888: 
1889: 
1890: Finally we address the corrections for the case of a spin-$3/2$ particle represented by a Rarita-Schwinger
1891: field $\psi_\mu$. At $\cO(a^2)$ terms that break $O(4)$ appear in the Lagrangian as
1892: \begin{equation} \label{eq:3/2}
1893: \cL = 
1894: a^2 c_{3/2} \, \ol \psi_\nu \, v_\mu v_\mu v_\mu v_\mu \, \psi_\nu
1895: +
1896: a^2 \ol c_{3/2} \, \ol \psi_\mu \, v_\mu v_\mu \, \psi_\mu 
1897: .\end{equation}
1898: To deduce the corrections to the energy, we cast this Lagrangian into its RPI form. 
1899: Under a reparameterization, one must maintain the conditions $- i \rlap \slash v \psi_\mu = \psi_\mu$ and $v_\mu \psi_\mu = 0$. 
1900: The RPI field $\Psi_\mu$ is a combination of terms from RPI invariant spinor and vector fields and is given by
1901: \begin{equation}
1902: \Psi_\mu 
1903: = 
1904: \left[
1905: \d_{\mu \nu} - \frac{1}{2 M} 
1906: \left( 
1907: \Dslash {}_\perp + i v_\mu D_\nu
1908: \right) 
1909: - \frac{1}{8 M^2}
1910: \left( 
1911: D^2 - 4 i v_\mu \Dslash {}_\perp D_\nu
1912: - 8 D_\mu D_\nu 
1913: \right)
1914: \right] 
1915: \psi_\nu
1916: ,\end{equation}
1917: up to $\cO(M^{-2})$. Half of the terms in the RPI spin-$3/2$ field vanish on shell. The remaining terms 
1918: are identical to those for a spin-$1/2$ particle. Thus for an on-shell particle, we have
1919: $\ol \Psi_\mu \Psi_\mu = \ol \psi_\mu \psi_\mu + \cO(M^{-3})$. Thus the on-shell RPI Lagrangian
1920: corresponding to Eq.~\eqref{eq:3/2} is given by
1921: \begin{eqnarray}
1922: \cL &=& 
1923: a^2 c_{3/2} \, \ol \psi_\nu 
1924: \left(v_\mu + \frac{i D_\mu}{M}\right)
1925: \left(v_\mu + \frac{i D_\mu}{M}\right)
1926: \left(v_\mu + \frac{i D_\mu}{M}\right)
1927: \left(v_\mu + \frac{i D_\mu}{M}\right)
1928: \psi_\nu \notag \\
1929: && + 
1930: a^2 \ol c_{3/2} \, \ol \psi_\mu 
1931: \left(v_\mu + \frac{i D_\mu}{M}\right)
1932: \left(v_\mu + \frac{i D_\mu}{M}\right)
1933: \psi_\mu \label{eq:RS}
1934: .\end{eqnarray}
1935: As with the spin-one case, we can evaluate the corrections to the spin-$3/2$ energy in the non-relativistic limit
1936: using the Lagrangian Eq.~\eqref{eq:RS} and explicit forms of the Rarita-Schwinger vectors. The result for the $\D$ mass
1937: is presented in the main text, see Eq.~\eqref{eq:new}.
1938: 
1939: 
1940: \bibliography{hb}
1941: 
1942: \end{document}
1943: