spin dynamics in pyrochlore heisenberg antiferromagnets

4
Spin Dynamics in Pyrochlore Heisenberg Antiferromagnets P. H. Conlon * and J. T. Chalker Theoretical Physics, Oxford University, 1 Keble Road, Oxford, OX1 3NP, United Kingdom (Received 9 April 2009; published 12 June 2009) We study the low temperature dynamics of the classical Heisenberg antiferromagnet with nearest neighbor interaction on the frustrated pyrochlore lattice. We present extensive results for the wave vector and frequency dependence of the dynamical structure factor, obtained from simulations of the preces- sional dynamics. We also construct a solvable stochastic model for dynamics with conserved magneti- zation, which accurately reproduces most features of the precessional results. Spin correlations relax at a rate independent of the wave vector and proportional to the temperature. DOI: 10.1103/PhysRevLett.102.237206 PACS numbers: 75.40.Gb, 75.10.Hk, 75.40.Mg Geometrical frustration in magnets inhibits ordering. Simple, classical models for these systems have very de- generate ground states [1,2]. Reflecting this degeneracy, highly frustrated magnetic materials characteristically re- main in the paramagnetic phase even at temperatures low compared to the scale set by exchange interactions. The behavior in this cooperative paramagnetic regime has been the focus of much recent research [3,4]. Nearest neighbor antiferromagnets on the pyrochlore lattice with classical n-component spins are representative of a large class of models [5,6]. They have remarkable correlations at low temperature, which are intermediate between those of conventionally ordered and completely disordered systems. These can be understood by mapping spin states onto configurations of a vector field, or flux field, which is solenoidal for ground states [79]. Gaussian fluctuations of this flux field provide a coarse-grained description of the cooperative paramagnet. Static spin correlations have a power-law dependence on separation, inside a correlation length $ that diverges as temperature T approaches zero. These correlations result in sharp fea- tures, termed pinch points, in diffuse scattering as a func- tion of wave vector. The dynamics of cooperative paramagnets has not been studied as extensively as the statics, but some ingredients are clear. In a Heisenberg model with precessional dynam- ics, the short-time behavior can be viewed in terms of harmonic spin wave fluctuations in the vicinity of a specific ground state, while over longer times the system wanders around the ground-state manifold. This second component to the motion results in decay of the spin autocorrelation function at long times, with a decay rate shown to be linear in T using simulations and phenomenological arguments [5]. These theoretical ideas are supported by inelastic neutron scattering measurements on pyrochlore antiferro- magnets: an early study of CsNiCrF 6 revealed strong tem- perature dependence to the width in energy of quasielastic scattering for T< j CW j [10], while recent work on Y 2 Ru 2 O 7 shows a width linear in T as predicted [11]. Our aim in this Letter is to establish a much more comprehensive description of cooperative paramagnets with precessional dynamics than has been available so far. The topic is interesting from several perspectives. First, in view of the pinch points in static correlations, it is natural to ask about the wave vector dependence of the dynamical structure factor, accessible in single-crystal measurements. Little is currently known about this: the autocorrelation function of Ref. [5] is expressed as an integral over all wave vectors, while the measurements of Ref. [11] used a powder sample. Second, dynamics in the paramagnetic phase of unfrustrated antiferromagnets is dominated by spin diffusion [1214], and one would like to know whether this extends to the cooperative paramag- net. Third, behavior in the Heisenberg model should be compared to that in spin ice, which is represented by the Ising pyrochlore antiferromagnet with dynamics controlled by the motion of monopole excitations [15]. In outline, our results are as follows. We find at low temperature three types of behavior in different regions of reciprocal space. (i) Close to reciprocal lattice points, correlations are dominated by spin diffusion with a temperature-independent diffusion constant. (ii) At a ge- neric wave vector [not included in (i) or (iii)] correlations are Lorentzian in frequency with a width linear in T and independent of wave vector. (iii) Close to nodal lines in reciprocal space on which the static, ground-state structure factor vanishes [8], dynamical correlations are dominated by finite-frequency spin wave contributions. This picture is hence very different from that for the kagome Heisenberg antiferromagnet, which shows order-by-disorder and propagating modes [16]. We consider the classical Heisenberg antiferromagnet with nearest neighbor interactions on the pyrochlore lat- tice. Lattice sites (labeled i; j) form corner-sharing tetrahe- dra (labeled ; ). Spins S i are unit vectors and L ¼ P i2 S i is the total spin of tetrahedron . The Hamiltonian is H ¼ J X hiji S i S j 1 2 J X L 2 þ c; (1) where c is a constant. Ground states satisfy L ¼ 0 for all PRL 102, 237206 (2009) PHYSICAL REVIEW LETTERS week ending 12 JUNE 2009 0031-9007= 09=102(23)=237206(4) 237206-1 Ó 2009 The American Physical Society

Upload: j-t

Post on 06-Apr-2017

214 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Spin Dynamics in Pyrochlore Heisenberg Antiferromagnets

Spin Dynamics in Pyrochlore Heisenberg Antiferromagnets

P.H. Conlon* and J. T. Chalker

Theoretical Physics, Oxford University, 1 Keble Road, Oxford, OX1 3NP, United Kingdom(Received 9 April 2009; published 12 June 2009)

We study the low temperature dynamics of the classical Heisenberg antiferromagnet with nearest

neighbor interaction on the frustrated pyrochlore lattice. We present extensive results for the wave vector

and frequency dependence of the dynamical structure factor, obtained from simulations of the preces-

sional dynamics. We also construct a solvable stochastic model for dynamics with conserved magneti-

zation, which accurately reproduces most features of the precessional results. Spin correlations relax at a

rate independent of the wave vector and proportional to the temperature.

DOI: 10.1103/PhysRevLett.102.237206 PACS numbers: 75.40.Gb, 75.10.Hk, 75.40.Mg

Geometrical frustration in magnets inhibits ordering.Simple, classical models for these systems have very de-generate ground states [1,2]. Reflecting this degeneracy,highly frustrated magnetic materials characteristically re-main in the paramagnetic phase even at temperatures lowcompared to the scale set by exchange interactions. Thebehavior in this cooperative paramagnetic regime has beenthe focus of much recent research [3,4].

Nearest neighbor antiferromagnets on the pyrochlorelattice with classical n-component spins are representativeof a large class of models [5,6]. They have remarkablecorrelations at low temperature, which are intermediatebetween those of conventionally ordered and completelydisordered systems. These can be understood by mappingspin states onto configurations of a vector field, or flux field,which is solenoidal for ground states [7–9]. Gaussianfluctuations of this flux field provide a coarse-graineddescription of the cooperative paramagnet. Static spincorrelations have a power-law dependence on separation,inside a correlation length � that diverges as temperature Tapproaches zero. These correlations result in sharp fea-tures, termed pinch points, in diffuse scattering as a func-tion of wave vector.

The dynamics of cooperative paramagnets has not beenstudied as extensively as the statics, but some ingredientsare clear. In a Heisenberg model with precessional dynam-ics, the short-time behavior can be viewed in terms ofharmonic spin wave fluctuations in the vicinity of a specificground state, while over longer times the system wandersaround the ground-state manifold. This second componentto the motion results in decay of the spin autocorrelationfunction at long times, with a decay rate shown to be linearin T using simulations and phenomenological arguments[5]. These theoretical ideas are supported by inelasticneutron scattering measurements on pyrochlore antiferro-magnets: an early study of CsNiCrF6 revealed strong tem-perature dependence to the width in energy of quasielasticscattering for T < j�CWj [10], while recent work onY2Ru2O7 shows a width linear in T as predicted [11].

Our aim in this Letter is to establish a much morecomprehensive description of cooperative paramagnets

with precessional dynamics than has been available sofar. The topic is interesting from several perspectives.First, in view of the pinch points in static correlations, itis natural to ask about the wave vector dependence of thedynamical structure factor, accessible in single-crystalmeasurements. Little is currently known about this: theautocorrelation function of Ref. [5] is expressed as anintegral over all wave vectors, while the measurements ofRef. [11] used a powder sample. Second, dynamics in theparamagnetic phase of unfrustrated antiferromagnets isdominated by spin diffusion [12–14], and one would liketo know whether this extends to the cooperative paramag-net. Third, behavior in the Heisenberg model should becompared to that in spin ice, which is represented by theIsing pyrochlore antiferromagnet with dynamics controlledby the motion of monopole excitations [15].In outline, our results are as follows. We find at low

temperature three types of behavior in different regions ofreciprocal space. (i) Close to reciprocal lattice points,correlations are dominated by spin diffusion with atemperature-independent diffusion constant. (ii) At a ge-neric wave vector [not included in (i) or (iii)] correlationsare Lorentzian in frequency with a width linear in T andindependent of wave vector. (iii) Close to nodal lines inreciprocal space on which the static, ground-state structurefactor vanishes [8], dynamical correlations are dominatedby finite-frequency spin wave contributions. This picture ishence very different from that for the kagome Heisenbergantiferromagnet, which shows order-by-disorder andpropagating modes [16].We consider the classical Heisenberg antiferromagnet

with nearest neighbor interactions on the pyrochlore lat-tice. Lattice sites (labeled i; j) form corner-sharing tetrahe-dra (labeled �;�). Spins Si are unit vectors andL� ¼ P

i2�Si is the total spin of tetrahedron �. TheHamiltonian is

H ¼ JXhiji

Si � Sj � 1

2JX�

L2� þ c; (1)

where c is a constant. Ground states satisfy L� ¼ 0 for all

PRL 102, 237206 (2009) P HY S I CA L R EV I EW LE T T E R Sweek ending12 JUNE 2009

0031-9007=09=102(23)=237206(4) 237206-1 � 2009 The American Physical Society

Page 2: Spin Dynamics in Pyrochlore Heisenberg Antiferromagnets

�. The equation of motion, describing precession of eachspin around its local exchange field, is

dSi

dt¼ �JSi �

Xj

Sj; (2)

where sites j are the nearest neighbors of i. The global spinrotation symmetry of Eq. (1) implies conservation of totalspin in the dynamics.

Before presenting results from a molecular dynamicsstudy of Eq. (2), we consider an analytically tractablestochastic model for the dynamic behavior. It is knownthat static spin correlators for the classical Heisenbergmodel are well described by those for n-component spinsin the large n limit [6,7]. Building on this, we set out toendow the n ¼ 1 model with appropriate dynamics. Firstwe recall some details of the static model. Taking thesecond form of the Hamiltonian in Eq. (1), a single spincomponent in the large n limit has the unnormalizedprobability distribution e��E with

�E ¼ 1

2

Xi

�s2i þ1

2�J

X�

l2�; (3)

where now l� ¼ Pi2�si is the sum of ‘‘soft’’ spins (�1<

si <1) on tetrahedron �. The spin length is constrainedby the Lagrange multiplier �. For � ! 1, the second termin Eq. (3) enforces all the l� to be zero. The interactionterm written directly in terms of the spins is 1

2�JP

ijðAij þ2�ijÞsisj, where Aij is the adjacency matrix for the pyro-

chlore lattice. We call the combination Aij þ 2�ij the

interaction matrix. Its eigenvalues v�ðqÞ are labeled by

wave vector q and a band index� 2 f1; 2; 3; 4g. Two bandsare flat [v1;2ðqÞ ¼ 0] and two (� ¼ 3; 4) are dispersive.

Requiring hs2i i ¼ 1=3 to mimic behavior of a single spincomponent in the Heisenberg model, � ¼ 3=2þOðT=JÞfor T � J. We denote the Fourier transform of the spinvariables si by s

aq, where a is a sublattice index, and define

the sublattice sum sq ¼ P4a¼1 s

aq. Transforming from saq to

the basis (denoted by tildes) that diagonalizes the interac-tion matrix gives collective spin variables ~s�q . We want to

introduce time dependence and calculate the dynamic cor-relation function Sðq; tÞ ¼ hsqðtÞs�qð0Þi, and its time

Fourier transform, the dynamic structure factor Sðq; !Þ,measured using neutron scattering.

There are many choices of dynamics which reproduceany given equilibrium distribution. To approximate Eq. (2)we demand a local dynamics that conserves the total spin.We can ensure this by requiring the spin on each site tosatisfy a local continuity equation. We introduce spincurrents on bonds of the pyrochlore lattice, which havedrift and noise terms. We take the drift current on a bondlinking two sites to be proportional to the difference in thegeneralized forces @E=@si at the sites. This favors relaxa-tion towards a configuration that minimizes E; the thermalensemble is maintained by noise which has an independent

Gaussian distribution on each bond. These assumptionslead to the dynamical equations for the soft spins

dsidt

¼ �Xl

�il

@E

@slþ �iðtÞ; (4)

where the matrix � is the lattice Laplacian (for a latticewith coordination number z, �il ¼ Ail � z�il). The corre-lator of the noise �iðtÞ at site i, h�iðtÞ�jðt0Þi ¼ 2T��ij�ðt�t0Þ, has an amplitude fixed by the requirement of thermalequilibrium. The only free parameter in the model is therate �, which sets a time scale for dynamical processes.The Langevin equation (4) is straightforward to solve inthe diagonal basis. It gives the correlation function

h~s�q ðtÞ~s��qð0Þi ¼���T

Jv� þ �Te��ð8�v�ÞðJv�þ�TÞt: (5)

The dynamic correlation function is then

Sðq; tÞ ¼ X4�¼1

g�ðqÞh~s�q ðtÞ~s��qð0Þi; (6)

where the structure factors g�ðqÞ are formed from the

eigenvectors of the interaction matrix. They satisfy thesum rule

P4�¼1 g�ðqÞ ¼ 4.

For completeness we present their explicit forms here. In

the notation of [7], where cab ¼ cosðqaþqb4 Þ and cab ¼

cosðqa�qb4 Þ with Q ¼ c2xy þ c2xy þ c2yz þ c2yz þ c2xz þ c2xz �

3 and defining P � ffiffiffiffiffiffiffiffiffiffiffiffiffi1þQ

p, the eigenvalues of the inter-

action matrix are v1;2 ¼ 0 and v3;4 ¼ 4� 2P. Further

defining s2a � sin2ðqa4 Þ and cðabÞ � cab þ cab, the g�ðqÞare for the degenerate flat bands

g � g1 þ g2 ¼ 2� 4

3�Q½cðxyÞs2z þ cðyzÞs2x þ cðzxÞs2y�

and for the dispersive bands

g3;4 ¼ 2� 1

2gð1 2P�1Þ P�1ð2� cðyzÞ � cðxzÞ � cðxyÞÞ

which indeed satisfy g1 þ g2 þ g3 þ g4 ¼ 4.We now examine the implications of this model for T �

J, emphasizing the features (i)–(iii) mentioned in our in-troduction. From the exponent in Eq. (5) we obtain acharacteristic time for decay of correlations. (i) In thevicinity of q ¼ 0 only the coefficient g4ðqÞ is nonzeroand so behavior is controlled by the fourth band whosedecay rate is �1 ¼ 8�Ja2q2 þOðq4Þ, where a is thepyrochlore site spacing; from this we identify the spindiffusion constant in this model as D ¼ 8�Ja2, indepen-dent of T. (ii) At a generic wave vector where g1 and g2 arenonzero, most of the spectral weight is in the flat bands,with decay rate �1 ¼ 8��T, independent of q; in anapproximation where only the flat bands contribute, thisimplies the dynamic structure factor factorizes asSðq; !Þ ¼ SðqÞfð!Þ, a possibility noted in [6]. (iii) On

PRL 102, 237206 (2009) P HY S I CA L R EV I EW LE T T E R Sweek ending12 JUNE 2009

237206-2

Page 3: Spin Dynamics in Pyrochlore Heisenberg Antiferromagnets

nodal lines [8], high symmetry directions in reciprocalspace along which g1ðqÞ þ g2ðqÞ ¼ 0, the decay rate iswave vector dependent and OðJÞ away from q ¼ 0.

To test these ideas we have performed simulations of thefull precessional dynamics, Eq. (2). Low T configurationsare obtained using a Metropolis Monte Carlo samplingmethod. We take these as initial configurations for numeri-cal integration of the equations of motion using a fourth-order Runge-Kutta algorithm with adaptive step size.Energy and total spin are conserved to relative errors nogreater than 10�6. We report data from simulations onsystem sizes with total number of sites N ¼ 4L3 for L ¼16; 32. We calculate the dynamic correlation functionSðq; tÞ � hSqðtÞ � S�qð0Þi and the dynamical structure fac-

tor Sðq; !Þ � hjSqð!Þj2i, where Sq ¼ P4a¼1 S

aq is the

Fourier transformed spin configuration and a runs overthe four sublattices. We present results under the headings(i)–(iii) as above, expressing q in reciprocal lattice units asq ¼ 2k for Figs. 2–4.

(i) Since the total magnetization is conserved, one ex-pects diffusion at sufficiently small q. The simulationsconfirm diffusive behavior with a diffusion constant inde-pendent of temperature. At small q, the data should col-lapse onto the scaling form appropriate for diffusion,

�q2Sðq; !Þ ¼ 3�2D

ð!=q2Þ2 þD2; (7)

where � is the susceptibility per primitive unit cell and thefactor of 3 is due to the three spin components. Figure 1shows this scaling collapse when plotted as in Eq. (7), anddemonstrates that the diffusion constant is independent oftemperature. The prediction of Eq. (6), whose small q limithas precisely the form of Eq. (7), is an excellent fit with� ¼ 0:167.

(ii) At a generic point in reciprocal space (not near q ¼0 or a nodal line) the structure factor is well described by aLorentzian centered on ! ¼ 0, indicating relaxational dy-namics (Fig. 2, upper inset). The decay rate for this relaxa-tion (Fig. 2, main panel) is proportional to T andindependent of the wave vector, even close to the pinchpoints.(iii) On nodal lines, by contrast, the width in frequency

of Sðq; !Þ isOðJÞ and depends little on T; see Fig. 2, lowerinset. High frequency (! ¼ 2:5J) and zero frequency be-havior is also presented in Fig. 3, as a survey of Sðq; !Þ inthe ðqx; qx; qzÞ plane, using data taken at �J ¼ 500 andtypical of all low temperatures. Weight in Sðq; !Þ at! Jand low T can be viewed as due to spin wave fluctuations inthe vicinity of an instantaneous ground state. In contrast tobehavior in the kagome antiferromagnet [16], there is noevidence in Fig. 3 for sharp propagating modes.We next demonstrate that the stochastic model accu-

rately reproduces most aspects of the precessional dynam-ics. We examine behavior with each type of dynamics atdifferent temperatures and times along a path in the

Brillouin zone: P ¼ ð0; 0; 0Þ !P 1 ð2; 2; 2Þ !P 2 ð0; 0; 2Þ. Thisconsists of the section P 1 along high symmetry nodal linespassing through a pinch point, and a section P 2 typical ofreciprocal space. We compare in Fig. 4 the dynamic corre-lation function, normalized to unity at t ¼ 0, at two tem-peratures with the predictions from Eq. (6) along the pathP . The excellent agreement of the curves across multipletemperatures, wave vectors, and times is good evidencethat the stochastic model is sufficient to capture the relaxa-tion behavior. It fails only at short times (t & J�1, notshown in Fig. 4) on nodal lines, where it does not accountfor oscillatory spin wave contributions.

10−3 10−2 10−110−3

10−2

10−1

−0.1 0 0.10

500

(a)

0 1 2 30

0.05

0.125

(b)

FIG. 2 (color online). Main panel: Decay rate of Sðq; tÞ as afunction of T at wave vectors as indicated. Upper inset: Sðq; !Þat �J ¼ 40 (red curve) and 100 (blue curve), for k ¼ ð1; 1; 1Þ.Lower inset: Sðq; !Þ on a nodal line at k ¼ ð0; 0; 1:25Þ.

0.20.4

0.60.8

−2 −1 0 1 2

0

25

50

75

100

125

(a)

−1 0 10

5

10

15

20 (b)

FIG. 1 (color online). Evidence for spin diffusion. (a) De-pendence of Sðq;!Þ on q and ! at small q. (b) Scaling collapsefollowing Eq. (7) at multiple temperatures (�J ¼ 20; 40; 60; 80)and four of the wave vectors plotted in (a). Also plotted is theprediction of Eq. (6).

PRL 102, 237206 (2009) P HY S I CA L R EV I EW LE T T E R Sweek ending12 JUNE 2009

237206-3

Page 4: Spin Dynamics in Pyrochlore Heisenberg Antiferromagnets

The stochastic model is microscopic and its main ingre-dients are a conservation law and the pyrochlore latticestructure. A long wavelength description is provided by themapping to flux fields [7,8] and it is interesting to see howthe dynamics translates under this mapping. Taking the lowtemperature, small q limit, the correlators for the contin-uum flux fields Bðq; tÞ implied by the stochastic model are

hBiðq; tÞBjð�q;0Þi /��ij �

qiqj

q2

�e�8��Tt

þ�qiqj

q2� qiqj

q2 þ��2

�e�8�ðJa2q2þ�TÞt:

This result can be derived from a Langevin equation for thecontinuum flux fields in which the ‘‘monopole density’’� ¼ r �B obeys a continuity equation @t�þr � j ¼ 0,with monopole current density

j ¼ 8��TB� 8�Ja2r�þ ðtÞ: (8)

Here, the second term is the usual diffusion current arisingfrom a density gradient, while the first describes responseto an entropic force. This response involves a drift currentof the magnetic charge density � in the field B that mimicselectrical conduction in electrodynamics and is responsiblefor the flat relaxation rate. Related results have been ob-tained recently in a study of dynamics in spin ice, repre-sented by the Ising antiferromagnet [15]. In this casemonopoles are discrete and it has been argued that purelydiffusive dynamics are insufficient to explain observationsand a full description must include the network of Diracstrings between monopoles, which are essentially entropic,as well as dipolar interactions [15]. In spin ice, dipolarinteractions lead to Coulomb-law forces between mono-poles. By contrast, in Eq. (8) Coulombic forces appearpurely entropically.

In summary, we have considered wave vector and fre-quency resolved dynamics of the classical pyrochlore anti-ferromagnet. The relaxational behavior is well captured bya stochastic model that conserves total spin. Spin diffuseswith a diffusion constant independent of temperature, andentropic forces drive currents to relax configurations with arate independent of wave vector and inversely proportionalto temperature.This work was supported in part by EPSRC Grant

No. EP/D050952/1.

*[email protected][1] P.W. Anderson, Phys. Rev. 102, 1008 (1956).[2] J. Villain, Z. Phys. B 33, 31 (1979).[3] A. P. Ramirez, Annu. Rev. Mater. Sci. 24, 453 (1994).[4] R. Moessner and A. P. Ramirez, Phys. Today 59, No. 2, 24

(2006).[5] R. Moessner and J. T. Chalker, Phys. Rev. B 58, 12 049

(1998); Phys. Rev. Lett. 80, 2929 (1998).[6] B. Canals and D.A. Garanin, Can. J. Phys. 79, 1323

(2001).[7] S. V. Isakov, K. Gregor, R. Moessner, and S. L. Sondhi,

Phys. Rev. Lett. 93, 167204 (2004).[8] C. L. Henley, Phys. Rev. B 71, 014424 (2005).[9] D. A. Huse, W. Krauth, R. Moessner, and S. L. Sondhi,

Phys. Rev. Lett. 91, 167004 (2003).[10] M. J. Harris, M. P. Zinkin, and T. Zeiske, Phys. Rev. B 52,

R707 (1995).[11] J. van Duijn, N. Hur, J.W. Taylor, Y. Qiu, Q. Z. Huang,

S.-W. Cheong, C. Broholm, and T. G. Perring, Phys. Rev.B 77, 020405(R) (2008).

[12] P. G. de Gennes, J. Phys. Chem. Solids 4, 223 (1958).[13] M. de Leener and P. Resibois, Phys. Rev. 152, 318 (1966).[14] A. Bunker, K. Chen, and D. P. Landau, Phys. Rev. B 54,

9259 (1996).[15] L. Jaubert and P. Holdsworth, Nature Phys. 5, 258 (2009).[16] J. Robert, B. Canals, V. Simonet, and R. Ballou, Phys. Rev.

Lett. 101, 117207 (2008).

(0,0,0) (1,1,1) (2,2,2) (0,0,2)

0

1

FIG. 4 (color online). Solid blue lines: Normalized correla-tion function Sðq; tÞ=Sðq; 0Þ. Dashed red lines: Prediction fromEq. (6) with � ¼ 0:167, both shown at 3 times, t.

−1 0 1 2 3−1

01

23

0

2.5

(a)

0 1 2 3 4 5 6 7 8

(0,0,0) (1,1,1)0

1

2

3

4

(b)

FIG. 3 (color online). Intensity map of �Sðq; !Þ in theðqx; qx; qzÞ plane at �J ¼ 500. (a) Lower panel: Zero frequency(data divided by 1:25� 105); white line is path P . Upperpanel: ! ¼ 2:5J; black circle is centered on a pinch point.(b) Section along path segment P 1 (see text).

PRL 102, 237206 (2009) P HY S I CA L R EV I EW LE T T E R Sweek ending12 JUNE 2009

237206-4