1: \documentclass[12pt]{article}
2: %\documentstyle[psfig,12pt]{article}
3:
4: \textheight=8.57truein
5: \textwidth=6.1truein
6: \overfullrule=0pt
7: \parskip=2pt
8: \parindent=12pt
9: \headheight=0in
10: \headsep=0in
11: \topmargin=0.3in
12: \oddsidemargin=0in
13:
14: %--------+---------+---------+---------+---------+---------+---------+
15: \newsavebox{\ns}
16: \newsavebox{\dbrane}
17: \newsavebox{\dbshort}
18:
19: \renewcommand{\arraystretch}{1.2}
20:
21: \usepackage{epsfig}
22: \usepackage{amssymb}
23:
24: %\newcommand{\sect}[1]{\setcounter{equation}{0}\section{#1}}
25: %\renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
26: %\def\appendix{{\newpage\section*{Appendix}}\let\appendix\section%
27: % {\setcounter{section}{0}
28: % \gdef\thesection{\Alph{section}}}\section}
29:
30: \def\be{\begin{eqnarray}}
31: \def\ee{\end{eqnarray}}
32:
33:
34: \newcommand{\nn}{\nonumber}
35: \newcommand\para{\paragraph{}}
36: \newcommand{\ft}[2]{{\textstyle\frac{#1}{#2}}}
37: \newcommand{\eqn}[1]{(\ref{#1})}
38: \newcommand\balpha{\mbox{\boldmath $\alpha$}}
39: \newcommand\bbeta{\mbox{\boldmath $\beta$}}
40: \newcommand\bgamma{\mbox{\boldmath $\gamma$}}
41: \newcommand\bomega{\mbox{\boldmath $\omega$}}
42: \newcommand\blambda{\mbox{\boldmath $\lambda$}}
43: \newcommand\bmu{\mbox{\boldmath $\mu$}}
44: \newcommand\bphi{\mbox{\boldmath $\phi$}}
45: \newcommand\bzeta{\mbox{\boldmath $\zeta$}}
46: \newcommand\bsigma{\mbox{\boldmath $\sigma$}}
47: \newcommand\bepsilon{\mbox{\boldmath $\epsilon$}}
48: \newcommand\btau{\mbox{\boldmath $\tau$}}
49:
50: \def\Dslash{\,\,{\raise.15ex\hbox{/}\mkern-12mu D}}
51: \def\Dbarslash{\,\,{\raise.15ex\hbox{/}\mkern-12mu {\bar D}}}
52: \def\delslash{\,\,{\raise.15ex\hbox{/}\mkern-9mu \partial}}
53: \def\delbarslash{\,\,{\raise.15ex\hbox{/}\mkern-9mu {\bar\partial}}}
54: \def\pslash{\,\,{\raise.15ex\hbox{/}\mkern-9mu p}}
55: \def\calDslash{\,\,{\raise.15ex\hbox{/}\mkern-12mu {\cal D}}}
56: \newcommand\Bprime{B${}^\prime$}
57: \newcommand{\sign}{{\rm sign}}
58:
59:
60: \def\pp{\mathbf{\Phi}}
61: \def\be{\begin{eqnarray}}
62: \def\ee{\end{eqnarray}}
63: \def\tr{{\rm Tr}}
64: \def\dd{{\rm d}}
65: \def\D{{\cal D}}
66: \def\vac{|\,0\rangle}
67: \def\nn{\nonumber}
68: % bold face in formulas
69: \def\mbf#1{\mathchoice{\hbox{\boldmath $\displaystyle #1$}}
70: {\hbox{\boldmath $\textstyle #1$}}
71: {\hbox{\boldmath $\scriptstyle #1$}}
72: {\hbox{\boldmath $\scriptscriptstyle #1$}}}
73:
74: \def\dblone{\hbox{$1\hskip -1.2pt\vrule depth 0pt height 1.6ex width 0.7pt
75: \vrule depth 0pt height 0.3pt width 0.12em$}}
76:
77:
78: \def\bR{\mathbb{R}}
79: \def\bC{\mathbb{C}}
80: \def\bH{\mathbb{H}}
81: \def\bO{\mathbb{O}}
82: \def\bF{\mathbb{F}}
83: \def\bP{\mathbb{P}}
84: \def\bK{\mathbb{K}}
85:
86:
87:
88: \begin{document}
89: \pagestyle{plain}
90: \setcounter{page}{1}
91: \newcounter{bean}
92: \baselineskip16pt
93:
94:
95: \begin{titlepage}
96:
97: \begin{center}
98: \today
99: \hfill {.}\\
100: %\hfill MIT-CTP-3390 \\
101:
102:
103: \vskip 1.5 cm {\large \bf The Berry Phase of D0-Branes}
104: \vskip 1cm
105: {Chris Pedder, Julian Sonner and David Tong}\\
106: \vskip 1cm {\sl Department of Applied Mathematics and Theoretical
107: Physics, \\ University of Cambridge, UK}
108:
109: \end{center}
110:
111:
112: \vskip 0.5 cm
113: \begin{abstract}
114: We study $SU(2)$ Yang-Mills quantum mechanics with $N=2,4,8$ and
115: $16$ supercharges. This describes the non-relativistic dynamics of
116: a pair of D0-branes moving in $d=3,4,6$ and $10$ spacetime
117: dimensions respectively. We show that as the D0-branes orbit,
118: states undergo a Berry holonomy described by the four Hopf maps. For the $N=2$
119: theory, the associated Hopf map is the ${\bf Z}_2$ M\"obius bundle
120: and its effect is to turn the D0-branes into anyons with exchange
121: statistics $e^{i\pi/2}$. For the $N=4,8$ and $16$ theories, the
122: Hopf maps give rise to Berry connections that are familiar to
123: physicists: the $U(1)$ Dirac monopole; the $SU(2)$ Yang monopole;
124: and the $SO(8)$ octonionic monopole.
125:
126:
127:
128: \end{abstract}
129:
130:
131: \end{titlepage}
132:
133: \subsection*{1. Introduction}
134:
135: Many years ago, Kugo and Townsend \cite{kugo} pointed out a
136: relationship between supersymmetric field theories with $N=2,4,8$
137: and $16$ supercharges and the four normed division algebras
138: $\bK\cong \bR$, $\bC$, $\bH$ and $\bO$. The key observation is
139: algebraic. Theories with $N=2,4,8$ and $16$ supercharges naturally
140: live in $d=3,4,6$ and $10$ spacetime dimensions respectively. The
141: Lorentz Lie algebra in these dimensions is isomorphic to the
142: algebra of $2\times 2$ Hermitian matrices with vanishing trace and
143: elements in $\bK$,
144: %
145: \be sl(2;\bK) \cong so(d-1,1) \label{the}\ee
146: %
147: This allows us to express a spinor in $d$ dimensions as a
148: 2-component $\bK$-vector, generalizing the well-known result for
149: $d=4$. This construction was elaborated upon in \cite{jon}.
150:
151: \para
152: For theories with $N=2,4$ and $8$ supercharges, the relationship
153: to the division algebra $\bK$ also manifests itself in more
154: physical and dynamical matters. This includes familiar features of
155: supersymmetric theories, such as the holomorphy of the
156: superpotential and the hyperK\"ahler/quaternionic structure of
157: Calabi-Yau moduli spaces. However, so far the tantalizing idea
158: that an octonionic structure underlies theories with 16
159: supercharges has not led to major insight about quantum dynamics.
160:
161: \para
162: The purpose of this short note is to point out that the
163: isomorphism \eqn{the} has a simple, physical consequence in
164: the framework of supersymmetric quantum mechanics. We focus on
165: $SU(2)$ gauged quantum mechanics. The theory with $N$ supercharges
166: can be thought of as the dimensional reduction of the minimal
167: super Yang-Mills theory in $d=3,4,6$ and 10 dimensions, and
168: describes the non-relativistic dynamics of two D0-branes moving in
169: $d-1=N/2+1$ spatial dimensions. We show that the states of a pair
170: of orbiting D0-branes undergo a holonomy described by the Hopf map
171: associated to the division algebra $\bR$, $\bC$, $\bH$ and $\bO$.
172:
173: \para
174: The four Hopf maps take $S^{N-1}\rightarrow S^{N/2}$:
175: %
176: \be \begin{array}{c} S^{1} \\ \ \,\Bigg{\downarrow} {}^{S^0\cong Z_2} \\
177: S^1
178: \end{array}\ \ \ \ \begin{array}{c} S^{3} \\ \ \,\Bigg{\downarrow} {}^{S^1} \\
179: S^2
180: \end{array}\ \ \ \ \begin{array}{c} S^{7} \\ \ \,\Bigg{\downarrow} {}^{S^3} \\
181: S^4
182: \end{array}\ \ \ \ \begin{array}{c} S^{15} \\ \ \,\Bigg{\downarrow} {}^{S^7} \\
183: S^8
184: \end{array}\nn\ee
185: %
186: %
187: %In each case, the total manifold $S^{N-1}$ can be thought of as
188: %arising from the pair $(q_1,q_2)\in \bK^2$, subject to the
189: %constraint $|q_1|^2 + |q_2|^2=1$. Meanwhile, the base manifold is
190: %the projective space, $S^{N/2}\cong \bK\bP^1$. We will describe
191: %the Hopf map in more detail in Section 3.
192: %
193: The case of $N=2$ supercharges is special. The map ${\bf
194: Z}_2\hookrightarrow S^1\rightarrow S^1$, sometimes known as the
195: zeroth Hopf map, describes the M\"obius bundle. We will show in
196: Section 3.1 that the ground state of the associated quantum
197: mechanics undergoes a discrete ${\bf Z}_2$ holonomy, changing by a
198: sign as D0-branes orbit. Since a single orbit is equivalent to two
199: exchanges, this implies that the D0-branes in $N=2$ Yang-Mills
200: quantum mechanics are anyons: they obey exchange statistics
201: $e^{\pm i \pi/2}=\pm i$. As we review in Section 3, the remaining
202: three Hopf maps induce connections over the base space $S^{d-2}$
203: which subsequently arise as Berry connections in the quantum
204: mechanics. Each is familiar to physicists: they are the $U(1)$
205: Dirac monopole, the $SU(2)$ Yang monopole \cite{yang}, and the
206: $SO(8)$ octonionic monopole \cite{grossman,tchrak}.
207:
208: \para
209: This quartet of Hopf bundles are the fab four of geometrical
210: phases. The discrete ${\bf Z}_2$ holonomy was one of the earliest
211: known Berry phases \cite{chemist}, while the Dirac monopole
212: appeared as an example in Berry's original paper \cite{berry}.
213: Both the Yang monopole and the octonionic monopole have also
214: arisen in various contexts. The former was first introduced as a
215: non-Abelian Berry phase in \cite{avron,feher}, and has subsequently
216: appeared in several condensed matter systems
217: \cite{bigzhang,middlezhang,smallzhang}, with various properties
218: explored in \cite{levay}. The $SO(8)$ octonionic monopole appeared
219: previously in association with the eight-dimensional quantum Hall
220: effect \cite{bern}. In this paper we will see how all these
221: connections arise as (non-Abelian) Berry phases in Yang-Mills
222: quantum mechanics with $N$ supercharges.
223:
224: \para
225: The present paper can be thought of as a follow-up to our recent
226: work \cite{us,second} exploring the Berry phase that emerges in
227: various supersymmetric quantum mechanics. In \cite{us}, we showed
228: that certain $N=4$ quantum systems naturally give rise to the
229: Dirac monopole and associated constructs such as the smooth 't
230: Hooft-Polyakov monopole. In \cite{second}, we studied a quantum
231: theory with $N=8$ supercharges and showed that a deformation of
232: the Yang-monopole arises as the spin connection of a dual
233: gravitational background. The Yang monopole and octonionic monopole
234: also appeared in a related context in \cite{usp}.
235:
236:
237:
238:
239:
240:
241:
242: \subsection*{2. D0-Brane Dynamics}
243:
244: In this paper we study the $SU(2)$ super-Yang-Mills quantum
245: mechanics with $N=2,4,8$ and $16$ supercharges. The Lagrangians
246: are the dimensional reduction of minimal super Yang-Mills in
247: $d=3,4,8$ and $10$ dimensions respectively, and each takes the
248: form
249: %
250: \be \label{lag} L = \frac{1}{2g^2}\,\tr \left( \left( \D_0 X_i
251: \right)^2 + \sum_{i<j}\left[ X_i , X_j \right]^2 +i \bar\psi\D_0
252: \psi + \bar\psi \Gamma^i \left[ X_i , \psi \right]
253: \right)\,, \ee %
254: where $X_i$ and $\psi_\alpha$ are both $su(2)$ valued fields. The
255: inverse coupling constant $1/g^2$ is the mass of the D0-brane.
256: Lagrangians with different amounts of supersymmetry are
257: distinguished by the number of scalar fields $X_i$, and the
258: structure of the Grassmann parameters $\psi$. The scalar index
259: runs over $i=1,\ldots, d-1$ and the Lagrangian describes the
260: non-relativistic relative motion of a pair of D0-branes moving in
261: $d-1=2,3,5$ and $9$ spatial dimensions\footnote{More precisely,
262: the theory with $N=16$ supercharges describes the dynamics of a
263: pair of D0-branes in flat ten-dimensional spacetime. Lagrangians
264: with less supersymmetry describe the dynamics of a pair of
265: fractional D0-branes, trapped to lie at a suitable singularity in
266: a K3, a CY 3-fold, or a $G_2$ holonomy manifold.}. The matrices
267: $\Gamma^i$ satisfy the $d-1$ dimensional Euclidean Clifford
268: algebra: $\{\Gamma^i,\Gamma^j\}=2\delta^{ij}$. The spinors are
269: real or complex, depending on the possible representations of the
270: Clifford algebra:
271: %
272: \begin{itemize}
273: \item $N=2$: For $d-1=2$, we may choose the real Pauli matrices
274: $\Gamma^1=\sigma^1$, $\Gamma^2=\sigma^3$. The spinor
275: $\psi_\alpha$, with $\alpha=1,2$, is also real.
276: %
277: \item $N=4$: For $d-1=3$, there are only complex representations
278: of the Clifford algebra. We take the Pauli matrices
279: $\Gamma^i=\sigma^i$. The spinor $\psi_\alpha$, with $\alpha=1,2$,
280: is complex.
281: %
282: \item $N=8$: The Clifford algebra in $d-1=5$ dimensions is again
283: complex. The complex spinor $\psi_\alpha$ has $\alpha=1,\ldots,4$.
284: %
285: \item $N=16$: The $d-1=9$ Clifford algebra admits a real
286: representation. The real spinor $\psi_\alpha$ has
287: $\alpha=1,\ldots,16$.
288: \end{itemize}
289: %
290:
291:
292: \subsubsection*{2.1 The Born-Oppenheimer Approximation}
293:
294: We work in the Born-Oppenheimer approximation, considering
295: well-localized wavepackets describing D0-branes with separation
296: $\hat{X}^i$. To this end, we expand around the classical
297: background
298: %
299: \be \langle X^i\rangle = \ft12 \hat{X}^i\sigma^3 \label{bo}\ee
300: %
301: and study the effects of high-frequency modes of open strings
302: stretched between the D0-branes. Our results are applicable in the
303: weak coupling limit $\hat{X}^3\gg g^2$. This typically excludes
304: the ground state of the full system, for which the wavefunction is
305: localized close to the origin $\hat{X} \sim g^{2/3}$.
306:
307: \para
308: The background \eqn{bo} breaks the gauge symmetry to the Cartan
309: subalgebra: $SU(2)\rightarrow U(1)$. The remnant ${\bf Z}_2$ Weyl
310: group acts as $\hat{X}^i\rightarrow -\hat{X}^i$. The configuration
311: space of the D0-branes is thus ${\bf R}^d/{\bf Z}_2$, reflecting
312: the indistinguishability of the particles. We expand,
313: %
314: \be \label{bmasses} X^i = \langle X^i\rangle + \ft12 x^i_m
315: \sigma^m \ee
316: %
317: The bosonic commutator term in \eqn{lag} provides oscillation
318: frequencies for the fluctuations $x^i_m$. The fluctuations
319: $x^i_3$, lying in the unbroken $U(1)$, are zero modes and describe
320: the relative motion of the D0-brane pair in $\bR^{d-1}$. In
321: contrast, the $m=1,2$ components, lying in the broken part of the
322: gauge group, pick up non-zero frequencies and describe the
323: excitation of open strings stretched between the D0-branes. We
324: form the complex combination
325: %
326: \be z^i=x^i_1+ix^i_2\ ,\ \ \ \ i=1,\ldots, d-1\ee
327: %
328: which has charge $+1$ under the unbroken $U(1)$ gauge symmetry.
329: Introducing the complex conjugate momentum $\pi_i$, the leading
330: order free Hamiltonian for these off-diagonal bosonic modes is
331: given by,
332: %
333: \be H_B = \frac{1}{2g^2}\left[ {\cal P}_{ij}\left( \pi_i
334: \bar\pi_j+ \hat X^2 z_i\bar z_j \right) + \ldots
335: %{\cal O}\left(\frac{g}{|\hat{X}|^{1/3}}\right)
336: \right] \ee
337: %
338: where $\ldots$ are interaction parts of the Hamiltonian which are
339: perturbations of order $g/\hat{X}^{1/3}$. The Hamiltonian includes
340: the projection operator
341: %
342: \be {\cal P}_{ij} = \delta_{ij}-\frac{\hat X^i \hat X^j}{\hat
343: X^2}\label{bproj}\ee
344: %
345: and therefore describes only $d-2$ complex harmonic oscillators.
346: These oscillators can be thought of as the transverse excitations
347: of a string stretched between the D0-branes. The remaining degree
348: of freedom falls victim to the broken gauge symmetry on the
349: D0-branes; it is analogous to the scalar that is eaten by the
350: Higgs mechanism in higher dimensions.
351:
352: \para
353: We may also expand the fermions around the background \eqn{bo}.
354: Once again, the fermions $(\psi_\alpha)_3$, lying in the Cartan
355: subalgebra, provide zero modes. Upon quantization, these fill out
356: a $2^{N/2}$ dimensional multiplet of states whose degeneracy is
357: split only by interaction terms. In contrast, the off-diagonal
358: components $(\psi_\alpha)_1$ and $(\psi_\alpha)_2$ have non-zero
359: frequencies. Here it is convenient to differentiate between the
360: $N=2,16$ cases, which have real fermions and the $N=4,8$ cases,
361: which have complex fermions. In the former case, we introduce the
362: complexified Grassmann parameter,
363: %
364: \be \label{eq:complexspinor} \Psi=\psi_1 + i \psi_2\,. \ee
365: %
366: in terms of which the fermionic Hamiltonian is written as
367: %
368: \be H_F^{N=2,16} = \frac{1}{2g^2}\left[\Psi^\dagger \left( \hat X
369: \cdot \Gamma \right)\Psi + \ldots\right]\label{hf216}\ee
370: %
371: The $N=4,8$ cases have complex spinors from the outset. We may now
372: form two linearly independent complex combinations
373: %
374: \be \Psi = \frac{1}{\sqrt{2}} \left( \psi_1 + i \psi_2 \right)\
375: \ \ ,\ \ \ \tilde{\Psi} = \frac{1}{\sqrt{2}} \left( \psi_1 - i
376: \psi_2 \right) \ee
377: %
378: with respective charges $+1$ and $-1$ under the unbroken $U(1)$
379: gauge group. The fermionic part of the Hamiltonian is now given by
380: %
381: \be H_F^{N=4,8}=\frac{1}{2g^2}\left[\Psi^\dagger \left(\hat X
382: \cdot \Gamma \right)\Psi - \tilde{\Psi}^\dagger \left( \hat X
383: \cdot \Gamma \right)\tilde{\Psi}+\ldots\right] \label{hf48}\ee
384: %
385: To summarize, the free part of the Hamiltonian $H=H_B+H_F$ for the
386: massive oscillators contains $N/2$ complex scalars and $N$ complex
387: Grassmann parameters. The interaction Hamiltonian is of order
388: $g/\hat{X}^{1/3}$.
389:
390:
391: \subsubsection*{2.2 Quantization and the Hilbert Space}
392:
393: In the Born-Oppenheimer approximation, we treat the massive
394: oscillator states $z_i$, $\Psi_\alpha$ and $\tilde{\Psi}_\alpha$
395: quantum mechanically in the classical background $\hat{X}$. The
396: zero frequency modes $\hat{X}^i$ and $(\psi_\alpha)_3$ are
397: quantized subsequently. The approximation holds as long as
398: $\hat{X}\gg g^{1/3}$ ensuring that we keep a separation of scales,
399: meaning that the wavefunction for $\hat{X}$ should not have
400: significant support near the origin where the two D0-branes
401: approach. For example, this will be the case for excited states of
402: orbiting D0-branes that carry large angular momentum. In the
403: following section we will study the holonomy of the excited states
404: of the massive oscillators as the two D0-branes orbit. We first
405: briefly describe the Hilbert space of these states.
406:
407: \para
408: Working in the regime $\hat{X}\gg g^{1/3}$, we may restrict
409: attention to the free theory. The $d-2$ massive complex scalars
410: have ground state energy $E_B=(d-2)\hat{X}/2g^2$. One may
411: construct the Hilbert space by acting with (suitably projected)
412: creation operators ${\cal P}_{ij}\,a^\dagger_j$ and ${\cal
413: P}_{ij}\,\bar{a}^\dagger_j$, where
414: %
415: \be a^\dagger_i=\frac{(\hat{X}z_i
416: -i\bar{\pi}_i)}{\sqrt{2\hat{X}}}\ \ \ ,\ \ \ \bar{a}_i^\dagger =
417: \frac{(\hat{X} \bar{z}_i-i\pi_i)}{\sqrt{2\hat{X}}} \ee
418: %
419: create quanta of charge $+1$ and $-1$ respectively under the
420: unbroken $U(1)$ gauge symmetry. Canonical quantization for
421: fermions gives the brackets $\{ \Psi_\alpha,\Psi_{\beta}^\dagger
422: \} = \delta_{\alpha\beta}$ and
423: $\{\tilde{\Psi}_\alpha,\tilde{\Psi}_{\beta}^\dagger \} =
424: \delta_{\alpha\beta}$. We build the fermionic Hilbert space ${\cal
425: H}_F$ by picking a reference state $\vac$, satisfying
426: %
427: \be \Psi_\alpha\vac=\tilde{\Psi}_\alpha\vac=0\label{ref}\ee
428: %
429: We may then act upon $\vac$ with $\Psi^\dagger$ in the $N=2,16$
430: theories, and $\Psi^\dagger$ and $\tilde{\Psi}^\dagger$ in the
431: $N=4,8$ theories, to construct a Hilbert space of dimension
432: $\dim({\cal H}_F)=2^N$.
433:
434: \para
435: The free fermionic Hamiltonians \eqn{hf216} and \eqn{hf48} have a
436: unique ground state $|\Omega\rangle$ with energy
437: $E_F=-N\hat{X}/4g^2$. This ensures that the ground state energy of
438: the full theory is $E_0=E_B+E_F=0$. The fermionic ground state
439: always lies in the sector with half of the fermions excited. To
440: describe it, we first introduce the fermionic projection operators
441: %
442: \be P_\pm = \frac{1}{2}\left( 1 \pm \frac{\hat X\cdot
443: \Gamma}{|\hat X|}\right)\label{fproj}\ee
444: %
445: Then, for the $N=2,16$ theories, the (un-normalized) ground state
446: is given schematically by,
447: %
448: \be\label{eq:ground} |\Omega\rangle = \left(P_-
449: \Psi^\dagger\right)^{N/2} \vac \label{vac1}\ee
450: %
451: while, for the $N=4,8$ theories, the (un-normalized) ground state
452: is given by
453: %
454: \be |\Omega\rangle = \left(P_- \Psi^\dagger\right)^{N/4} \left(P_+
455: \tilde{\Psi}^\dagger\right)^{N/4} \vac \label{vac2}\ee
456: %
457: Note that $|\Omega\rangle$ is the ground state for the high
458: frequency modes in the Born-Oppenheimer approximation. The
459: question of whether a normalizable ground state of the full theory
460: exists is more subtle, and irrelevant for our considerations. (It
461: does for $N=16$, but is expected not to for $N=2,4,8$).
462:
463: \para
464: The physical Hilbert space of the theory is subject to Gauss' law
465: for the unbroken $U(1)\subset SU(2)$ gauge symmetry which ensures
466: that all physical states are gauge neutral. The vacuum
467: $|\Omega\rangle$ is assigned charge zero under the $U(1)$ and so
468: survives Gauss' purge. Other states in the fermionic Hilbert space
469: arise by acting on $|\Omega\rangle$ with combinations of the
470: charge $+1$ states $P_-\Psi$ and $P_-\tilde{\Psi}^\dagger$ and the
471: charge $-1$ states $P_+\Psi^\dagger$ and $P_+\tilde{\Psi}$. In
472: each case, one may construct a state in the physical Hilbert space
473: by dressing the fermionic state with appropriate powers of the
474: charged bosonic operator ${\cal P}_{ij}a_j^\dagger$ and ${\cal
475: P}_{ij}\bar{a}_j^\dagger$.
476:
477:
478:
479: \subsection*{3. Berry Phase and Hopf Maps}
480:
481: In the previous section we constructed the Hilbert space over each
482: background separation $\hat{X}^i$. In this section we are
483: interested in how these Hilbert spaces evolve as the D0-branes
484: orbit, and $\hat{X}^i$ traces a closed path $\Gamma$ in
485: configuration space. The dynamical phase of a set of degenerate
486: states $|\phi_a\rangle$ is accompanied by a (possibly non-Abelian)
487: Berry holonomy, described by
488: %
489: \be |\phi_a\rangle \longrightarrow P \exp\left(-i \oint_\Gamma
490: {A}_{ab}\cdot \hat{X} \right)|\phi_b\rangle \ \ \ {\rm with}\ \ \
491: (A_i)_{ab}=i\langle\phi_b|\frac{\partial}{\partial\hat{X}^i}|\phi_a\rangle\ee
492: %
493: We will see that for certain states in quantum mechanics with $N$ supercharges, the
494: Berry connection $A$ is given by the associated Hopf map. We start
495: by reviewing the the Hopf maps and connections in more detail.
496:
497: \para
498: The Hopf map from $S^{N-1}\rightarrow S^{N/2}$ is defined in the
499: following manner. One starts with a commuting spinor
500: $\chi_\alpha$, of the type described in Section 2: real for
501: $N=2,16$ and complex for $N=4,8$. Imposing the normalization
502: condition $\chi^\dagger \chi =1$ ensures that $\chi$ defines a
503: point on $S^{N-1}$. The map to the base manifold $S^{N/2}$ is then
504: given by the bi-linear form
505: %
506: %
507: \be n^i = \chi^\dagger \Gamma^i \chi \ee
508: %
509: where $\Gamma^i$ obey the $SO(d-1)$ Clifford algebra. One may
510: check that $n^in^i=1$ and hence $n^i$ defines a point on
511: $S^{N/2}$.
512:
513:
514: \para
515: It is illustrative to stress the connection to the division
516: algebras by reformulating the Hopf maps over the algebra $\bK$
517: using the isomorphism \eqn{the}. (For more details see, for example,
518: \cite{baez}). The total manifold $S^{N-1}$ can be thought of as
519: arising from the pair $(q_1,q_2)\in \bK^2$, subject to the
520: constraint $|q_1|^2 + |q_2|^2=1$. From this pair we can define the
521: vector $n^i$, $i=1\ldots,d-1$,
522: %
523: \be n^i = q^\dagger \Gamma^i q\ee
524: %
525: where $\Gamma^i$ still satisfy the $SO(d-1)$ Clifford algebra,
526: hence justifying their name, but can now be thought of as basis
527: elements of $\Gamma^i \in sl(2;\bK)$, defined by
528: %
529: \be \Gamma^i=\left(\begin{array}{cc} 0 & e_i \\ e_i^\star & 0
530: \end{array}\right)\ \ \ i=1,\ldots , N/2\ \ \ ,\ \ \
531: \Gamma^{d-1}=\left(\begin{array}{cc} 1 & 0 \\ 0 & -1
532: \end{array}\right)\label{gam}\ee
533: %
534: where $e_i$ are the generators of the division algebra $\bK$. The
535: fact that these matrices obey the Clifford algebra reflects the
536: isomorphism \eqn{the}. One can check once again that $n^in^i=1$,
537: ensuring that this defines a map to the base manifold
538: $S^{N/2}\cong \bK\bP^1$.
539:
540: \para
541: The zeroth Hopf map associated to $\bR$ is the M\"obius bundle
542: ${\bf Z}_2\hookrightarrow S^1\rightarrow
543: \bR\bP^1$. We will shortly see how this
544: arises in the theory with $N=2$ supercharges: the holonomy of the
545: ground state is a minus sign whose role is to render the D0-branes
546: anyonic. The remaining Hopf maps, associated to $\bC$, $\bH$ and
547: $\bO$, each define a connection $A$ over the base space $S^{N/2}$,
548: which describes how the base manifold is twisted inside the total
549: space. These connections are given in terms of the projection
550: operators $P_\pm$ defined in \eqn{fproj}, where we identify
551: $\hat{X}^i\equiv \hat{X}n^i$. Let $\lambda_a$ be the non-vanishing
552: orthonormal eigenvectors of $P_-$: i.e. $P_-\lambda_a=\lambda_a$.
553: The Hopf connection is defined by,
554: %
555: \be A_{ab} = i\lambda_b^\dagger\, d\lambda_a\label{hopfcon}\ee
556: %
557: For $N=4,8$ and $16$, the maximum value of the index $a$ is
558: $1,2,8$ and $A_{ab}$ is therefore a $U(1)$, $U(2)$ and $SO(8)$
559: connection respectively. In the following, we will describe these
560: connections in more detail and see how they arise as the Berry
561: phase of certain states in the quantum mechanics.
562:
563:
564:
565: \subsubsection*{3.1 $N=2$ and the Zeroth Hopf Map}
566:
567: For the $N=2$ case, the ground state wavefunction undergoes
568: a discrete holonomy as the D0-branes orbit. While the existence
569: of this phase follows on general grounds from the degenerate
570: nature of $H_F$ at the origin, it is instructive to review
571: explicitly how it occurs.
572:
573: \para
574: The fermionic Hilbert space ${\cal H}_F$ consists of four states
575: $\vac$, $\Psi^\dagger_\alpha\vac$ and
576: $\Psi^\dagger_1\Psi^\dagger_2\vac$. The ground state may be
577: expanded as a linear combination of the middle sector,
578: %
579: \be |\Omega\rangle=\lambda_\alpha \Psi^\dagger_\alpha\vac\ee
580: %
581: with $\lambda_\alpha$ the negative eigenvector of $\hat{X}\cdot
582: \Gamma$.
583: %They have energy
584: %%
585: %\be H_F\vac=H_F\Psi^\dagger_1\Psi^\dagger_2\vac=0\ \ \ {\rm and}\
586: %\ \ H_F\Psi_\alpha^\dagger\vac =
587: %\frac{1}{2g^2}(\hat{X}\cdot\Gamma)_{\alpha\beta}\Psi^\dagger_\beta\vac\ee
588: %%
589: We choose to work with the real basis of gamma matrices
590: $\Gamma^1=\sigma^1$ and $\Gamma^2=\sigma^3$ and introduce polar
591: coordinates
592: %
593: \be X^1=X\sin\theta\ \ \ ,\ \ \ X^2=X\cos\theta\ee
594: %
595: Then the groundstate eigenvector is
596: %
597: \be \vec{\lambda} =
598: \frac{1}{\sqrt{2+2\cos\theta}}\left(\begin{array}{c}-\sin\theta \\
599: 1+ \cos\theta \end{array}\right) \label{turn}\ee
600: %
601: where we have resolved the square-root sign ambiguity in favour of
602: the positive. The evolution of the eigenvector as $\theta$ is
603: adiabatically varied is shown in the figure. We see that as
604: $\theta$ varies from 0 to $2\pi$, $\vec{\lambda}$ returns to
605: $-\vec{\lambda}$. This is the manifestation of the zeroth Hopf
606: map: $S^1\rightarrow S^1$.
607:
608: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
609: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
610: \newcommand{\onefigurenocap}[1]{\begin{figure}[h]
611: \begin{center}\leavevmode\epsfbox{#1.eps}\end{center}
612: \end{figure}}
613: \newcommand{\onefigure}[2]{\begin{figure}[htbp]
614: \begin{center}\leavevmode\epsfbox{#1.eps}\end{center}
615: \caption{\small #2\label{#1}}
616: \end{figure}}
617: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
618: \begin{figure}[htbp]
619: \begin{center}
620: \epsfxsize=2in\leavevmode\epsfbox{z2berry.eps}
621: \end{center}
622: {\small Figure 1: The ${\bf Z}_2$ holonomy as the eigenvector
623: encircles the degeneracy at the origin.}
624: %\label{mon}
625: \end{figure}
626: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
627:
628: \para
629: This ${\bf Z}_2$ Berry phase is inherited by the ground state of
630: the D0-branes. After a single orbit, the wavefunction returns to
631: $|\,\Omega\rangle \mapsto -|\,\Omega\rangle$. Yet such an orbit is
632: equivalent to two exchanges of the D0-branes. Since the branes are
633: indistinguishable particles, upon a single exchange,
634: the D0-brane wavefunction picks up the phase $e^{\pm i\pi/2}=\pm
635: i$, where the $\pm$ sign depends on whether the exchange proceeds
636: clockwise or anti-clockwise. We learn that the D0-branes are
637: anyons \cite{anyon1,anyon2}.
638:
639: \para
640: It is worthwhile re-deriving the fact that the D0-branes are
641: anyons by studying a single exchange, rather than a complete
642: orbit. Let us start with the D0-branes separated by
643: $\hat{X}^i=(0,\hat{X})$, for which the ground state is given by
644: $|\Omega\rangle_+ = \Psi_2^\dagger\vac$. After an anti-clockwise
645: rotation to $\hat{X}^i=(0,-\hat{X})$, we see from \eqn{turn} that
646: the ground state adiabatically evolves into $|\Omega\rangle_- =
647: -\Psi_1^\dagger|0\rangle$. The ${\bf Z}_2$ Weyl symmetry ensures
648: that the Hilbert space constructed over $\hat{X}^i$ is physically
649: identified with the Hilbert space over $-\hat{X}^i$. Thus in order
650: to understand the phase picked up by the ground state, we need to
651: understand the map between the Hilbert spaces constructed over
652: $\pm\hat{X}^i$ induced by the Weyl group.
653:
654: \para
655: The action of the ${\bf Z}_2$ Weyl group is given by,
656: %
657: \be &{\bf Z}_2:& \hat{X}^i \mapsto -\hat{X}^i\ \ \ ;\ \ \
658: z^i\mapsto -z^{\dagger i}\ \ \ ;\ \ \ \Psi\mapsto -\Psi^\dagger
659: \ee
660: %
661: From the latter of these actions, we see that the reference state
662: $\vac_+$ defined by \eqn{ref} at $\hat{X}^i$ is mapped to
663: $\Psi_1^\dagger \Psi_2^\dagger \vac_-$ at $-\hat{X}^i$, up to a
664: phase $\omega$,
665: %
666: \be &{\bf Z}_2:& \vac_+ \mapsto \omega \Psi_1^\dagger
667: \Psi_2^\dagger \vac_-\ee
668: %
669: The question is: what is $\omega$? We can answer this by following the
670: fate of the state $\Psi_1^\dagger\vac_+$ under two actions of the
671: Weyl group,
672: %
673: \be \Psi_1^\dagger\vac_+\mapsto -\omega\Psi_2^\dagger\vac_-\mapsto
674: -\omega^2 \Psi_1^\dagger\vac_+\ee
675: %
676: Insisting that the Weyl group action squares to the unit operator,
677: we learn that $\omega=\pm i$. We now use this knowledge to map the
678: state $|\Omega\rangle_-$ above back to an element of the Hilbert
679: space over $\hat{X}^i=(0,\hat{X})$. We see,
680: %
681: \be |\Omega\rangle_+ \stackrel{\tiny{\rm
682: adiabatic}}{\longrightarrow} |\Omega\rangle_- \stackrel{\tiny{\rm
683: {\bf Z}_2}}{\longrightarrow}\ \mp i|\Omega\rangle_+\ee
684: %
685: The upshot is that the ground state of the two D0-branes indeed
686: picks up the phase factor $\pm i$ upon exchange.
687:
688: \para
689: Note that we have still to quantize the $U(1)\subset SU(2)$ Cartan
690: subalgebra. In particular, this includes the massless fermions
691: $(\psi_3)_\alpha$, which double the Hilbert space. The resulting
692: states have conjugate exchange statistics. If one state picks up a
693: phase $+i$ upon an anti-clockwise rotation, the other state picks
694: up a phase $-i$.
695:
696: \subsubsection*{3.2 $N=4$ and the First Hopf Map}
697:
698: We now turn to D0-branes moving in $d-1=3$ spatial dimensions.
699: As explained in Section 2, the Hilbert space is constructed by
700: acting with bosonic and fermionic creation operators on the
701: reference state $\vac$. Each of these creation operators is
702: accompanied by a projection operator: ${\cal P}$ for bosons,
703: defined in \eqn{bproj}, and $P_\pm$ for fermions, defined in
704: \eqn{fproj}. These two types of projection operators give rise to
705: two types of Berry holonomy:
706:
707: \para
708: The projection operators ${\cal P}$ restrict the bosonic
709: excitations to lie tangent to the sphere $S^2$ at fixed
710: $|\hat{X}|$. As we saw above, this can be traced to the
711: implementation of the $SU(2)$ gauge symmetry and, from the string
712: theory perspective, is the familiar statement that stretched
713: strings have only transverse excitations. The net result is that as the
714: D0-branes orbit, excited states that include bosonic excitations
715: undergo a Berry holonomy as tangent vectors on the sphere $S^2$.
716: This is the same kind of
717: non-Abelian holonomy that was described in the original
718: paper of Wilczek and Zee \cite{wz}.
719:
720: \para
721: Fermionic excitations are accompanied by the projection operators
722: $P_\pm$ \eqn{fproj}. To be concrete, let's focus on the operator
723: $P_-$. We introduce the normalized eigenvector
724: $P_-\lambda=\lambda$. Acting with the creation operator
725: $P_-\Psi^\dagger$ means creating the normalized state
726: $|\lambda\rangle =\lambda_\alpha\Psi_\alpha^\dagger\vac$. As
727: $\hat{X}^i$ varies adiabatically, this state undergoes a holonomy
728: described by the Berry connection,
729: %
730: \be A_i = i \langle \lambda | \frac{\partial}{\partial
731: \hat{X}^i}|\lambda\rangle\ee
732: %
733: which is the connection of the Hopf map \eqn{hopfcon}. An explicit
734: form of the connection requires a choice of gauge which, in this
735: context, means a chosen phase for the eigenvector $\lambda$. We
736: choose $\lambda_1\in \bR$, a choice which is valid everywhere
737: except along the positive $\hat{X}^3$ axis. In this gauge, the
738: Berry connection takes the familiar form of the Dirac monopole,
739: %
740: \be A_i =
741: \frac{-\hat{X}^j}{2\hat{X}(\hat{X}-\hat{X}^3)}\,\epsilon_{ij}
742: \qquad i = 1,2 \ \ \ \ ,\ \ \ \ A_3 =0 \label{dirac}\ee
743: %
744: The Hopf-Berry connection has first Chern class $-1$ over the
745: sphere $S^2$, meaning that the integral of the field strength
746: $F=dA$ yields,
747: %
748: \be c_1=\frac{1}{2\pi}\int_{S^2} F = -1\ee
749: %
750: Acting with the operator $P_+\Psi^\dagger$ results is a state
751: whose Berry connection has Chern class $+1$. The ground state in
752: the $N=4$ theory is given by
753: %
754: \be |\Omega\rangle = N^\prime(P_-\
755: \Psi^\dagger)(P_+\tilde{\Psi}^\dagger)\vac\equiv (\lambda_\alpha
756: \Psi_\alpha^\dagger)(\tilde{\lambda}_\beta\tilde{\Psi}_\beta^\dagger)\vac\ee
757: %
758: where $N^\prime$ is a normalization factor, while $\lambda$
759: ($\tilde{\lambda}$) is the normalized non-zero eigenvector of
760: $P_\mp$. The presence of the two, opposite, projection operators
761: ensures that the ground state does not pick up a Berry phase as
762: the D0-branes orbit.
763:
764: \para
765: Excited states do pick up a Berry phase, given by the sum of the
766: phases associated to the relevant projection operators. For
767: example, a degenerate pair of states obeying Gauss' law
768: is given
769: by
770: %
771: \be |\phi_-\rangle = (\lambda_\alpha\Psi_\alpha^\dagger)
772: (\lambda_\alpha\tilde{\Psi}_\alpha^\dagger)\vac \ \ \ ,\ \ \
773: |\phi_+\rangle = (\tilde{\lambda}_\alpha\Psi_\alpha^\dagger)
774: (\tilde{\lambda}_\alpha\tilde{\Psi}_\alpha^\dagger)\vac \ee
775: %
776: These states have energy $E=E_B+E_F=\hat{X}/g^2$ and describe two
777: strings attached to the D0-branes. (They are part of a triplet of
778: excited spin 1 states).
779: %
780: %transform with charge with charge $\pm 2$ under the unbroken
781: %$U(1)_R$ R-symmetry of the problem, reflecting the fact that they
782: %come from a spin 1 state in three dimensions. (The third state in
783: %the triplet, $|\phi_0\rangle =
784: %(P_+\Psi^\dagger)(P_-\tilde{\Psi}^\dagger)\vac$, has had it's
785: %energy split by the separation of the D0-branes).
786: %
787: From the discussion above, we see that these states pick up a
788: Berry phase arising from a magnetic monopole of charge $q=\pm2$.
789: Physically, this means that the D0-branes orbit as charged
790: particles as if in the presence of a magnetic monopole fixed at
791: their centre. In a semi-classical analysis, the orbits are no
792: longer restricted to lie on a plane, but rather lie on a cone with
793: opening angle $\cos\theta = -q/(2J+q)$, where $J$ is angular
794: momentum of the spinning D0-branes. The energy $E$ these rotating
795: states scales as $E^3\sim g^2 J(J+q)$.
796:
797:
798:
799: \subsubsection*{3.3 $N=8$ and the Quaternionic Hopf Map}
800:
801: The discussion for the $N=8$ theory is very similar to the $N=4$
802: theory above. The unique ground state of the system \eqn{vac2}
803: does not undergo a Berry phase.
804: %
805: %\be |\Omega\rangle \sim
806: %(P_-\Psi^\dagger)^2(P_+\tilde{\Psi}^\dagger)^2
807: %%(\lambda_{1\alpha}\Psi_\alpha^\dagger)(\lambda_{2\beta}\Psi_\beta^\dagger)
808: %%(\tilde{\lambda}_{1\gamma}\Psi_\gamma^\dagger)
809: %%(\tilde{\lambda}_{2\delta}\Psi_\delta^\dagger)
810: %\vac\ee
811: %%
812: %This state does not undergo a Berry phase. Further, excited states
813: %do undergo a Berry holonomy.
814: %
815: Excited states do. The typical state undergoes a non-Abelian
816: holonomy arising from the sum of bosonic and fermionic Berry
817: connections. Once again, bosonic excitations give rise to a
818: holonomy in which states transform as tangent vectors on $S^5$.
819: More interesting for the present discussion are the fermionic
820: excitations. The projection operator $P_-$ now has a pair of
821: orthonormal eigenvectors: $P_-\lambda_a=\lambda_a$, $a=1,2$. This
822: means that the resulting Berry holonomy of the states
823: $|\lambda_a\rangle= \lambda_{a\alpha}\Psi^\dagger_\alpha\vac$ is
824: described by a $U(2)$ connection,
825: %
826: \be (A_i)_{ab} = i\langle \lambda_a |\frac{\partial}{\partial
827: \hat{X}^i}|\lambda_b\rangle \ee
828: %
829: Explicit computation shows that this is actually an $SU(2)$
830: connection, known as the Yang-monopole \cite{yang}. A suitable
831: choice of gamma matrices and gauge, can be found in \cite{second}.
832: The resulting connection is,
833: %
834: \be (A_i)_{ab} =
835: \frac{-\hat{X}^j}{2\hat{X}(\hat{X}-\hat{X}^5)}\,\eta^m_{ij}\,\sigma^{m}_{ab}
836: \qquad i = 1,2,3,4 \ \ \ \ ,\ \ \ \ A_5 =0 \label{yang}\ee
837: %
838: where $\eta_{\mu\nu}^m$ are the self-dual $4\times 4$ 't Hooft
839: matrices and $\sigma_{ab}$ are the Pauli matrices. The Yang
840: monopole is perhaps more familiar when viewed as a connection
841: restricted to $S^4$, where it is simply the $SO(5)$ invariant
842: instanton satisfying $F=-{}^\star F$ with second Chern class,
843: %
844: \be c_2=\frac{1}{8\pi^2}\int_{S^4}\,{\rm tr}(F\wedge F)=-1 \ee
845: %
846: where the non-Abelian field strength is defined by
847: $F_{ij}=\partial_iA_j-\partial_jA_i-i[A_i,A_j]$. States
848: constructed from the projection operators $P_+$ undergo a Berry
849: holonomy associated with a Yang-monopole of Chern class $c_2=+1$.
850:
851: %\para
852: %Again, we can focus on the fermionic Berry connection by looking
853: %at the two pairs of states,
854: %%
855: %\be |\phi\rangle = N(P_+\Psi^\dagger)(P_-\Psi)|\Omega\rangle\ \ \
856: %,\ \ \ |\tilde{\phi}\rangle = \tilde{N}
857: %(P_-\tilde{\Psi}^\dagger)(P_+\tilde{\Psi})|\Omega\rangle\ee
858: %%
859: %Each of these states removes one fermion from the ground state
860: %with lower energy, and replaces it with an excited state. We have
861: %neglected indices, but these describe two pairs of states, all
862: %with energy $E=2\hat{X}$. The pair of states $|\phi\rangle$
863: %undergo holonomy from the Yang monopole; the states
864: %$|\tilde{\phi}\rangle$ undergo holonomy from the Yin monopole.
865: %Yang-monopoles. {\bf Question: Is this right? This seems to be the
866: %direct sum of connections that we had in the previous paper,
867: %rather than the direct product that we might have expected}.
868:
869:
870:
871:
872: \subsubsection*{3.4 $N=16$ and the Last Hopf Map}
873:
874: Details for the theory with $N=16$ supercharges are again similar
875: to those above: the vacuum \eqn{vac1} does not pick up a Berry
876: connection, while bosonic excitations transform as tangent vectors
877: on $S^8$. The fermionic excitations are associated to projection
878: operators $P_\pm$, which are now $16\times 16$ real matrices. We
879: once again define the $8$ orthonormal, real eigenvectors
880: $P_-\lambda_a=\lambda_a$, with $a=1,\ldots,8$. States in the
881: quantum mechanics involving fermions undergo a holonomy arising
882: from the $SO(8)$ Berry connection,
883: %
884: \be (A_i)_{ab} = i\langle \lambda_b |\frac{\partial}{\partial
885: \hat{X}^i}|\lambda_a\rangle\ee
886: %
887: with $|\lambda_a\rangle =
888: \lambda_{a\alpha}\Psi_\alpha^\dagger\vac$. We refer to the
889: resulting connection as the $SO(8)$ octonionic monopole. It was
890: constructed in \cite{grossman,tchrak}. A simple expression for the
891: connection can be found in \cite{bern}
892: %
893: \be (A_i)_{ab} =
894: \frac{-\hat{X}^j}{2\hat{X}(\hat{X}-\hat{X}^9)}\,\Sigma_{ij} \qquad
895: i = 1,\ldots,8 \ \ \ \ ,\ \ \ \ A_9 =0 \label{so8}\ee
896: %
897: where $\Sigma_{ij}$ are the 28 generators of $SO(8)$ Lie algebra,
898: defined in terms of the $\Gamma$ matrices. An explicit form can be
899: given if we choose a suitable representation for the generators of
900: the octonions $e_i$, $i=1\ldots 8$ in \eqn{gam}, in terms of
901: $8\times 8$ real matrices. We choose the unit matrix to correspond
902: to $e_8=1$. Then $\Sigma_{ij} = -\ft12 [e_i,e_j]$, for
903: $i=1,\ldots,7$ and $\Sigma_{i8}=e_i$ \cite{bern}. The non-Abelian
904: field strength restricted to $S^8$ satisfies the generalized
905: self-duality condition $F\wedge F = -{}^\star F\wedge F$, with
906: %
907: \be \frac{1}{4!(2\pi)^4}\int_{S^8}{\rm tr}(F\wedge F\wedge
908: F\wedge F) = -1\ee
909: %
910: %We can now see how this connection acts on the states of the
911: %theory. Once again, the ground state $|\Omega\rangle =
912: %(P_-\Psi^\dagger)^8\vac$ has no Berry phase. The simplest states
913: %which do pick up a Berry connection are the $64$ states with
914: %energy $E=2\hat{X}$, with purely fermionic excitations
915: %%
916: %\be |\phi\rangle = N(P_+\Psi^\dagger)(P_-\Psi)|\Omega\rangle\ee
917: %%
918: %These excited states, corresponding to two strings stretched
919: %between the D0-branes, undergo a non-Abelian holonomy given by the
920: %tensor product of the octonionic instanton and anti-instanton.
921: %
922: %
923: As in previous cases, states associated to $P_+$ undergo a Berry
924: holonomy with $c_4=1$.
925:
926: \para
927: It is intriguing to see the octonionic monopole appearing in the
928: Matrix theory for D0-branes in this fashion, although its physical
929: significance in M-theory remains obscure. Nonetheless, it is
930: tempting to speculate. Firstly, recall that there is a precedent
931: for the appearance of Berry's phase in Matrix theory: the membrane
932: feels the magnetic field of the five-brane through the appearance
933: of a Berry connection \cite{dberk}. In this case, the abstract
934: Berry connection is recast as a physical magnetic field. It would
935: be very interesting if a similar M-theory interpretation could be
936: given in the present case. We note in passing that a physical
937: octonionic monopole is conjectured to act as the end-point for an
938: open heterotic string \cite{pol}.
939:
940: \subsection*{Acknowledgments}
941:
942: We would like to thank Freddy Cachazo, Joe Polchinski, Savdeep
943: Sethi and Paul Townsend for useful discussions. J.S. gratefully
944: acknowledges the University of Chicago, Harvard and the Perimeter
945: Institute for hospitality and travel support while this paper was
946: being written up. D.T. thanks the Isaac Newton Institute for their
947: kind hospitality while this work was undertaken. C.P. is supported
948: by an EPSRC studentship. J.S. is supported by a research
949: fellowship from Trinity College, Cambridge. D.T. is supported by
950: the Royal Society.
951:
952:
953:
954: \begin{thebibliography}{99}
955:
956: \small
957: \parskip=0pt plus 2pt
958:
959: %\cite{Kugo:1982bn}
960: \bibitem{kugo}
961: T.~Kugo and P.~K.~Townsend,
962: {\it Supersymmetry And The Division Algebras,}
963: Nucl.\ Phys.\ B {\bf 221}, 357 (1983).
964: %%CITATION = NUPHA,B221,357;%%
965:
966: \bibitem{jon} J.~M.~Evans, ``{\it Supersymmetric Yang-Mills Theories
967: and Division Algebras}'', Nucl.\ Phys.\ B {\bf 298}, 92 (1988).
968: %%CITATION = NUPHA,B298,92;%%
969:
970: \bibitem{yang} C.~N.~Yang,
971: ``{\it Generalization Of Dirac's Monopole To SU(2) Gauge Fields},''
972: J.\ Math.\ Phys.\ {\bf 19}, 320 (1978).
973: %%CITATION = JMAPA,19,320;%%
974: %\cite{Demler:1998pm}
975:
976:
977: \bibitem{grossman}
978: B.~Grossman, T.~W.~Kephart and J.~D.~Stasheff,
979: ``{\it Solutions To Yang-Mills Field Equations In Eight-Dimensions And The Last
980: Hopf Map},''
981: Commun.\ Math.\ Phys.\ {\bf 96}, 431 (1984)
982: [Erratum-ibid.\ {\bf 100}, 311 (1985)].
983: %%CITATION = CMPHA,96,431;%%
984:
985: \bibitem{tchrak} D.~H.~Tchrakian,
986: ``{\it Spherically Symmetric Gauge Field Configurations With Finite Action In 4
987: P-Dimensions (P = Integer)},''
988: Phys.\ Lett.\ B {\bf 150}, 360 (1985).
989: %%CITATION = PHLTA,B150,360;%%
990:
991: \bibitem{chemist} G. Herzberg and H. C. Longuet-Higgins,
992: ``{\it Intersection of potential energy surfaces in polyatomic
993: molecules}", Discuss. Faraday Soc., 1963, 35, 77
994:
995: \bibitem{berry} M.~V.~Berry,
996: ``{\it Quantal phase factors accompanying adiabatic changes},''
997: Proc.\ Roy.\ Soc.\ Lond.\ A {\bf 392}, 45 (1984).
998: %%CITATION = PRSLA,A392,45;%%
999:
1000: \bibitem{avron} J.~E.~Avron, L.~Sadun, J.~Segert and B.~Simon,
1001: ``{\it Topological Invariants in Fermi Systems with Time-Reversal
1002: Invariance}", Phys.\ Rev.\ Lett.\ {\bf 61}, 1329 (1988); ``{\it
1003: Chern Numbers, Quaternions and Berry's Phases in Fermi
1004: Systems}", Commun. Math. Phys. {\bf 124} (1989) 595.
1005: %%CITATION = PRLTA,61,1329;%%
1006:
1007: \bibitem{feher} M.~G.~Benedict, L.~G.~Feher and Z.~Horvath,
1008: ``{\it Monopoles and Instantons from Berry's Phase},''
1009: J.\ Math.\ Phys.\ {\bf 30}, 1727 (1989).
1010: %%CITATION = JMAPA,30,1727;%%
1011:
1012: \bibitem{bigzhang}
1013: E.~Demler and S-C.~Zhang,
1014: ``{\it Non-Abelian holonomy of BCS and SDW quasiparticles,}"
1015: Annals Phys.\ {\bf 271} (1999) 83
1016: [arXiv:cond-mat/9805404].
1017: %%CITATION = APNYA,271,83;%%
1018:
1019: \bibitem{middlezhang} S. Murakami, N. Nagaosa, S-C. Zhang,
1020: ``{\it SU(2) Non-Abelian Holonomy and Dissipationless Spin Current
1021: in Semiconductors}", Phys. Rev. {\bf B69}, 235206 (2004),
1022: [arXiv:cond-mat/0310005v3]
1023:
1024: \bibitem{smallzhang} C-H. Chern, H-D. Chen, C. Wu, J-P. Hu, S-C.
1025: Zhang, ``{\it Non-abelian Berry's phase and Chern numbers in
1026: higher spin pairing condensates}", Phys. Rev. {\bf B69}, 214512
1027: (2004) [arXiv:cond-mat/0310089v2].
1028:
1029:
1030: %\cite{Levay:1990}
1031: \bibitem{levay}
1032: P. L\'evay,
1033: {\it Geometrical description of SU(2) Berry phases,}
1034: Phys.\ Rev.\ A \ {\bf 41} {1990} 2837;
1035: ``{\it Quaternionic Gauge Fields And The Geometric Phase},''
1036: J.\ Math.\ Phys.\ {\bf 32}, 2347 (1991);
1037: ``{\it NonAbelian
1038: Born-Oppenheimer electric gauge force and the natural metric on
1039: Hilbert subspaces},''
1040: Phys.\ Rev.\ A {\bf 45}, 1339 (1992);
1041: %%CITATION = PHRVA,A45,1339;%%
1042: %%CITATION = JMAPA,32,2347;%%
1043:
1044: \bibitem{bern}
1045: B.~A.~Bernevig, J.~p.~Hu, N.~Toumbas and S.~C.~Zhang,
1046: ``{\it The Eight Dimensional Quantum Hall Effect and the Octonions},''
1047: Phys.\ Rev.\ Lett.\ {\bf 91}, 236803 (2003)
1048: [arXiv:cond-mat/0306045].
1049: %%CITATION = PRLTA,91,236803;%%
1050:
1051: \bibitem{us} C.~Pedder, J.~Sonner and D.~Tong,
1052: ``{\it The Geometric Phase in Supersymmetric Quantum Mechanics},''
1053: arXiv:0709.0731 [hep-th].
1054: %%CITATION = ARXIV:0709.0731;%%
1055:
1056: %\cite{Pedder:2007wp}
1057: \bibitem{second}
1058: C.~Pedder, J.~Sonner and D.~Tong,
1059: ``{\it The Geometric Phase and Gravitational Precession of D-Branes},''
1060: arXiv:0709.2136 [hep-th].
1061: %%CITATION = ARXIV:0709.2136;%%
1062:
1063:
1064:
1065: \bibitem{usp} B.~Chen, H.~Itoyama and H.~Kihara,
1066: ``{\it Nonabelian Berry phase, Yang-Mills instanton and USp(2k) matrix model},''
1067: Mod.\ Phys.\ Lett.\ A {\bf 14}, 869 (1999)
1068: [arXiv:hep-th/9810237];
1069: ``{\it Nonabelian monopoles from matrices:
1070: Seeds of the spacetime structure},''
1071: Nucl.\ Phys.\ B {\bf 577}, 23 (2000)
1072: [arXiv:hep-th/9909075].
1073: %%CITATION = NUPHA,B577,23;%%
1074:
1075:
1076:
1077:
1078: \bibitem{baez}
1079: J.~C.~Baez,
1080: ``{\it The Octonions},''
1081: arXiv:math/0105155.
1082: %%CITATION = MATH/0105155;%
1083:
1084: \bibitem{anyon1} J.~M.~Leinaas and J.~Myrheim,
1085: ``{\it On the theory of identical particles},''
1086: Nuovo Cim.\ B {\bf 37} (1977) 1.
1087: %%CITATION = NUCIA,37B,1;%%
1088:
1089:
1090: \bibitem{anyon2} F.~Wilczek,
1091: ``{\it Quantum Mechanics Of Fractional Spin Particles},''
1092: Phys.\ Rev.\ Lett.\ {\bf 49}, 957 (1982).
1093: %%CITATION = PRLTA,49,957;%%
1094:
1095:
1096: \bibitem{wz} F.~Wilczek and A.~Zee,
1097: ``{\it Appearance Of Gauge Structure In Simple Dynamical Systems},''
1098: Phys.\ Rev.\ Lett.\ {\bf 52}, 2111 (1984).
1099: %%CITATION = PRLTA,52,2111;%%
1100:
1101:
1102: \bibitem{dberk} M.~Berkooz and M.~R.~Douglas,
1103: ``{\it Five-branes in M(atrix) theory},''
1104: Phys.\ Lett.\ B {\bf 395}, 196 (1997)
1105: [arXiv:hep-th/9610236].
1106: %%CITATION = PHLTA,B395,196;%%
1107:
1108: \bibitem{pol} J.~Polchinski,
1109: ``{\it Open heterotic strings},''
1110: JHEP {\bf 0609}, 082 (2006)
1111: [arXiv:hep-th/0510033].
1112: %%CITATION = JHEPA,0609,082;%%
1113:
1114:
1115:
1116: \end{thebibliography}
1117:
1118:
1119:
1120:
1121:
1122:
1123:
1124:
1125:
1126: \end{document}
1127:
1128:
1129:
1130: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1131: %\newcommand{\onefigure}[2]{\begin{figure}[htbp]
1132: %
1133: % \caption{\small #2\label{#1}(#1)}
1134: % \end{figure}}
1135: %\newcommand{\onefigurenocap}[1]{\begin{figure}[h]
1136: % \begin{center}\leavevmode\epsfbox{#1.eps}\end{center}
1137: % \end{figure}}
1138: %\renewcommand{\onefigure}[2]{\begin{figure}[htbp]
1139: % \begin{center}\leavevmode\epsfbox{#1.eps}\end{center}
1140: % \caption{\small #2\label{#1}}
1141: % \end{figure}}
1142: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1143:
1144: %\begin{figure}[htb]
1145: %\begin{center}
1146: %\epsfxsize=4in\leavevmode\epsfbox{mir.eps}
1147: %\end{center}
1148: %\caption{Mirror symmetry at various scales.}
1149: %\label{mir}
1150: %\end{figure}
1151: