IUCM96-002 Non Possum Comprimi Ergo Sum: Skyrmions and Edge States in the Quantum Hall Effect arXiv:cond-mat/9601144v1 30 Jan 1996 A.H. MacDonald Indiana University, Department of Physics, Bloomington, IN 47405, USA. (October 5, 2017) Abstract When the chemical potential of an electron system has a discontinuity at a density n∗ , the system is said to be incompressible and a finite energy is required to create mobile charges in the bulk of the system. The quantum Hall effect is associated with incompressibilities in a two-dimensional electron system that occur at magnetic-field dependent densities, n∗ (B). In these notes we discuss two aspects of the physics of quantum Hall systems that follow directly from this association. Typeset using REVTEX 1 I. INTRODUCTION The quantum Hall effect [1,2] is an anomaly that occurs in the transport properties of two-dimensional electron system (2DES) in the regime of strong perpendicular magnetic fields. At certain magnetic fields it is found that the voltage drop in the system in the direction of current flow, which is responsible for dissipation, vanishes at low temperatures. Our understanding of this transport anomaly is not absolutely complete, however there is fairly broad agreement that the effect can occur only when the electronic system has, in the absence of disorder, jumps in its chemical potential at certain densities (n∗ ), that depend on magnetic field. In Section II of these notes we specifically discuss the relationship between incompressibility and transport properties in a 2DES. The remaining sections discuss two aspects of the physics of quantum Hall systems that are direct consequences of this relationship. Quantum Hall systems can be ferromagnetic. Under appropriate circumstances the spinmoments of the electrons align spontaneously, i.e. in the absence of Zeeman coupling to a magnetic field. We will refer to the 2DES in this case as a quantum Hall ferromagnet. Quantum Hall ferromagnets provide a particularly simple example of a two-dimensional itinerant electron ferromagnet, and therefore represent an attractive target for theories of the quantum statistical mechanics of such systems [3–5]. They also have a number of interesting unique properties, the most striking of which is that their instantons carry an electronic charge. This property is an immediate consequence of incompressibility at a magnetic-field dependent density. In Section III of these notes we discuss the electrically charged instantons of quantum Hall ferromagnets and some of the observable consequences of their presence. Another consequence of incompressibility at a magnetic field dependent density is that quantum Hall systems necessarily have gapless excitations localized at their edges. The lowenergy physics of quantum Hall systems is therefore like that of a one-dimensional electron system which, because of time-reversal symmetry breaking by the magnetic field, can carry a current even in equilibrium. In Section IV we will discuss the physics of quantum Hall 2 edges, including their description in terms of the Luttinger liquid models appropriate to one-dimensional fermion systems. Section V contains some concluding remarks. These notes do not attempt a systematic introduction to the physics of the quantum Hall effect. Interested readers can consult a previous effort of mine [6] that is cited frequently below. If greater depth is desired, readers can consult one of the excellent books covering different aspects of the subject [2]. I have written previously at this level but at greater length on quantum Hall edge states [7]; parts of the present notes borrow from that text. II. INCOMPRESSIBILITY AND THE QUANTUM HALL EFFECT The thermodynamic compressibility of a system of interacting particles is proportional to the derivative of the chemical potential with respect to density. It can happen that at zero temperature the chemical potential has a discontinuity at a density n∗ : the energy to add a particle to the system (µ+ ) differs, at this density, from the energy to remove a particle from the system (µ− ). The system is then said to be incompressible. In an incompressible system a finite energy is required to create unbound positive and negative charges that are capable of carrying current through the bulk. The number of these free charges present in the system will have an activated temperature dependence and will vanish for T → 0. Incompressible systems are usually insulating. Paradoxically, as we explain below, incompressibility is precisely the condition required for the quantum Hall effect to occur. The twist is that in the case of the quantum Hall effect, the density n∗ at which the incompressibility occurs must depend on magnetic field. In my view, incompressibility at a magnetic-field-dependent density is the sine qua non of the quantum Hall effect. For non-interacting electrons, the single-particle energy spectrum of a 2DES in a magnetic field consists of Landau levels, separated by hωc ≡ heB/m∗ c energy gaps and with a ¯ ¯ macroscopic degeneracy Nφ = AB/Φ0 = A/(2πℓ2). (Here B is the magnetic field strength, Φ0 = hc/e is the magnetic flux quantum, A is the area of the system, and m∗ is the electron mass.) Both the energy gaps between Landau levels and the degeneracy of the Landau levels 3 are proportional to B. Chemical potential discontinuities occur whenever the density is an integral multiple of n∗ = B/Φ0 . We show below that this property requires the existence of the gapless edge excitations discussed in Section IV. It is conventional in discussing the quantum Hall effect to use a magnetic field dependent density unit by defining the Landau level filling factor ν ≡ n/n∗ = 2πℓ2 n. For non-interacting electrons, incompressibilities occur at integer filling factors. When interactions are included, incompressibilities can also occur at fractional filling factors and, as we will discuss, physical properties near integer filling factors can be qualitatively altered. The relationship between incompressibility and the transport anomalies that give the quantum Hall effect its name can be understood by the following argument [8]. Consider a 2DES at zero temperature, as illustrated in Fig. [1]. We consider the case in which the chemical potential lies in the ‘charge gap’; µ ∈ (µ− , µ+ ). We want to consider the change in the equilibrium local currents, present in the system because of the breaking of time-reversal-invariance by the magnetic field, when we make an infinitesimal change in the chemical potential, δµ. Because µ lies in the charge gap the change in the local current density anywhere in the bulk of the system must be zero. The current density can change, if it does anywhere, only at the edge of the system. It follows from charge conservation that, if there is a change in the current flowing along the edge of the system, it must be the same at any point along the edge. We can relate this change in current to the change in the orbital magnetization: δI = c δM. A (1) Eq. (1) is just the equation for the magnetic moment of a current loop. However, δM = ∂N ∂M |B δµ = |µ δµ. ∂µ ∂B (2) (N is the number of electrons.) The second equality in Eq. (2) follows from a Maxwell relation. Combining Eq. (1) and Eq. (2) we obtain the following result for the rate at which the equilibrium edge current changes with chemical potential when the chemical potential lies in a charge gap: 4 δI ∂n∗ =c |µ . δµ ∂B (3) The fact that δI/δµ = 0 implies that whenever the charge gap occurs at a density that depends on magnetic field, there must be gapless excitations at the edge of the system. Properties of these low-energy edge excitations are discussed in Section IV. Eq.( 3) is expected to apply to the edge states even when the chemical potential lies only in a mobility gap and not in a true gap, as illustrated schematically in Fig. [1]. A net current can be carried from source to drain across the system by changing the local chemical potentials only at the edges and having different chemical potentials along the two edges connecting source and drain. When bulk states are localized, the two edges and the bulk are effectively decoupled from each other. Eq. (3) then also applies to transport currents, relating the current carried from source to drain to the chemical potential difference between the two edges, equal to eVH where VH is the Hall voltage. There is no voltage drop along an edge since each edge is in local equilibrium and hence no dissipation inside the sample. Eq. (3) is commonly known as the Stˇeda-Widom formula [9]. In using this picture to explain r transport experiments in bulk systems it is necessary to claim that the transport current will be carried entirely at the edge of the system even when bulk states occur at the Fermi level, as long as these states are localized. There are difficulties with this argument as a complete explanation for all transport phenomena associated with the quantum Hall effect, but that is another story and we will not pursue it here. III. QUANTUM HALL FERROMAGNETS A. Energy Scales The quantum Hall regime is usually understood as the regime in which no qualitative change in physical properties results from mixing of Landau levels by either interactions or disorder. It is common in theoretical studies to truncate the Hilbert space to a single orbital Landau level and include mixing , if at all, only when making quantitative estimates for 5 comparison with experiment. The quantum Hall regime, then, assumes that the Landau level separation, hωc is larger than other energy scales of interest. On the other hand the ¯ ferromagnetic state is generally defined in terms of the properties of an electronic system in the absence of a magnetic field. The term Quantum Hall Ferromagnet appears to be an oxymoron. To understand why it is not only sensible but also of more than academic interest to add this category to our taxonomy of electronic states, it is necessary to consider the relevant energy scales for the case of the semiconductor systems in which 2DES’s are realized. For a free-electron system in a magnetic field, the Zeeman splitting of spin-levels gµB B and the Landau level separation hωc are identical, apart from small relativistic corrections. ¯ Electrons in states near the conduction band minimum of a semiconductor behave like free electrons [10] except that band effects renormalize the electron mass m∗ and the g-factor. In the case of the GaAs systems, where the quantum Hall effect is most often studied, band effects increase the Landau level separation by a factor of ∼ 20 and reduce the Zeeman splitting by a factor of ∼ 4. As a result for typical experimental situations, the Landau level separation (in temperature units) is ≈ 200K, and the characteristic scale for electronelectron interactions is ≈ 100K while the Zeeman splitting is only ≈ 2K. We call a system a quantum Hall ferromagnet if the electronic spins in the incompressible ground state with density n∗ align in the absence of Zeeman coupling. In many cases, the properties of a quantum Hall ferromagnet with such a small Zeeman coupling do not differ noticeably from the properties when the Zeeman coupling is set to zero. In other cases, the small Zeeman coupling plays an important role but it is still useful to treat the system as a ferromagnet in the presence of a small symmetry breaking field. B. Ferromagnetic Ground States The Hartree-Fock approximation, in which many-electron states are approximated by single Slater determinants, provides a simple explanation for itinerant electron ferromagnetism that is often qualitatively correct. For non-interacting electrons, energy is minimized 6 by occupying both spin-states of each single particle energy level. (If the number of majorityspin electrons exceeds the number of minority-spin electrons it is necessary to occupy higher energy single-particle levels.) The ground state thus has equal numbers of majority-spin and minority-spin electrons if the total number of electrons is even and the difference is one if the total number of electrons is odd. This statement applies for an arbitrary spinquantization axis. As long as there is no spin-orbit coupling, the Hamiltonian is invariant under global spin rotations and the total spin of all electrons is a good quantum number. The above statement is equivalent to the observation that for non-interacting electrons the ground state always has total spin S = 0 if the number of electrons is even and total spin S = 1/2 if the number of electrons is odd. However, interaction energies are lower in singleSlater-determinant states with higher values of the total spin and, generally, is minimized in fully spin-polarized states with S = N/2. As in the familiar Hund’s rules from atomic physics, higher spin states tend to have lower interaction energies because like-spin electrons are prevented from being at the same position by the Pauli exclusion principle and therefore have more energetically favorable spatial correlations. In the Hartree-Fock approximation, or in closely related spin-density-functional approximations, an itinerant electron system is expected to be ferromagnetic if the reduction in interaction energy due to creating a finite spin-polarization state exceeds the cost in single-particle kinetic energy, or more generally ‘band’ energy. Because of Landau level degeneracy, the cost in kinetic energy of creating a finite spin polarization for electrons in a magnetic field is precisely zero unless ν is an even integer. Hartree-Fock or similar approximations would predict a ferromagnetic ground state for electrons at nearly any value of ν. In fact, this conclusion is incorrect. For example, at certain filling factors it is known [11] that the interaction energy is minimized in a S = 0 state. We do believe, however, that there exist finite ranges of filling factor over which the ground state has S/N = 0. For a deeper understanding of this behavior, we need a more rigorous argument. The approach we now describe is in the same spirit as the illuminating outlook on the spin-polarized fractional quantum Hall effect that arises from appropriate hard-core model 7 Hamiltonians [12,13]. As discussed for the case of interest below, these models have zero energy many-particle eigenstates that are often known analytically, are separated from other many-particle states by a finite gap, and have a degeneracy that increases with decreasing N. The incompressible state responsible [6] for a quantum Hall effect transport anomaly in such a model is the nondegenerate maximum N zero energy eigenstate. The zero energy eigenstates at lower densities constitute the portion of the spectrum that involves only the degrees of freedom of the, in general fractionally charged [14], quasiholes of the incompressible state. It is assumed that the difference between the model Hamiltonian and the true Hamiltonian is a sufficiently weak perturbation that the quasihole states are still well separated from other states in the Hilbert space, although accidental degeneracies will be lifted in the spectrum of the true Hamiltonian. Here we apply this approach to argue that the ground state at ν = 1 is a quantum Hall ferromagnet with S = N/2. For our analysis we use the symmetric gauge in which the single-particle orbitals [6] in the lowest Landau level are φm (z) = zm exp(−|z|2 /4), m+1 πm!)1/2 (2 (4) where [15] m = 0, 1, · · · , Nφ − 1, z = x + iy, and x and y are the Cartesian components of the two-dimensional coordinate. We study here a hard-core model for which the interaction is: V = 4πV0 i Nφ and θ = 0 for N < Nφ . |N − Nφ | is the number of Skyrmions or antiskyrmions present in the system. The integer quantum number K will depend in general on the relative size of Zeeman and Coulomb interaction terms and is the relevant quantum measure of the Skyrmion size. For non-interacting electrons, or with interactions treated in the Hartree-Fock approximation, K = 0 so that Sz always has the maximum value allowed by the Pauli exclusion principle. (K is guaranteed by particle-hole symmetry [26] to have the same value for N > Nφ and N < Nφ .) The NLσ model considerations of Sondhi et al. [18] described above, and also earlier numerical exact diagonalization calculations [27], suggest that K should be non-zero for quantum Hall ferromagnets and quite large if the Zeeman energy is small. These predictions were dramatically confirmed when Barrett et al. unexpectedly succeeded [28] in using optical pumping techniques to perform NMR Knight shift measurements of the spin-polarization of two-dimensional electron systems in the quantum Hall regime. The results of this experiment are illustrated in Fig. 3 and correspond to K = 3, in quantitative agreement with microscopic predictions based on a generalized Hartree-Fock approximation for single-Skyrmion states [29]. There seems to be little doubt that the elementary charged excitations of quantum Hall ferromagnets are Skyrmion-like objects that carry large spin quantum numbers. Recent transport [30] and optical [31] experiments add additional support to this conclusion. For large enough |N − Nφ | the Skyrmion-like objects will eventually interact strongly. When the density of Skyrmions is low and the temperature is low, Skyrmions are expected to form a lattice similar to the Wigner crystal state formed by electrons in the limit of very strong magnetic fields. In Fig. 4 we compare theoretical calculations of the spin-polarizations as a function of filling factor for several candidate Skyrmion lattice states with experimental data. The theoretical results were obtained by Brey et al. using a generalized Hartree-Fock approximation [32] and illustrate several important aspects of the physics of Skyrme crystals. These authors find that the ground state of the Skyrme crystal is a square lattice rather than 13 a triangular lattice as found for the electron Wigner crystal. Furthermore, as illustrated in Fig. 4, the spin-polarization of the square lattice Skyrme crystal is much smaller, for a given Zeeman coupling strength, than for the lowest energy triangular lattice state. The preference for a square lattice can be understood qualitatively in terms of the NLσ model description of Skyrmion states. For that model Skyrmions are centered at an arbitrary point, have an arbitrary size, and are invariant under arbitrary global spin rotations. When Coulomb and Zeeman energies are included the optimal Skyrmion size is fixed and the spin moment must be aligned with the Zeeman field far from the Skyrmion center. However, the energy of each Skyrmion is still invariant under global rotations of the moment about an axis aligned with the Zeeman field. For a Skyrme lattice the relative values of these rotation angles must be adjusted to minimize the total energy. It turns out that the interaction energy between a pair of Skyrmions is reduced when they have opposing orientations for the component of the ordered moment perpendicular to the Zeeman field. This arrangement allows the ordered moment orientation to vary more smoothly along the line connecting Skyrmion centers. The tendency toward opposing orientations is frustrated on triangular lattice, hence the energetic preference for a square lattice. The stronger short-range repulsive interaction in the aligned orientation ferromagnetic lattice case, results in smaller Skyrmions and therefore more spin-polarized states. The spin-polarizations calculated for the opposing orientation, square lattice case shown in Fig. 4 appear to be in excellent agreement with experiment over a wide range of filling factors near ν = 1. IV. EDGE EXCITATIONS OF AN INCOMPRESSIBLE QUANTUM HALL FLUID A. Non-Interacting Electron Picture Throughout this section we will consider a disk geometry where electrons are confined to a finite area centered on the origin by a circularly symmetric confining potential, Vconf (r). We have in mind the situation where Vconf (r) rises from zero to a large value near r = R, where R 14 is loosely speaking the radius of the disk in which the electron system is confined. We choose this geometry, for which the electron system has a single edge, since we limit our attention here to the properties of an isolated quantum Hall edge and will not discuss the physics of interaction or scattering between edges [33]. In this geometry it is convenient to choose the symmetric gauge for which angular momentum is a good quantum number. For Vconf (r) the single-electron kinetic energy operator has the macroscopically degenerate Landau levels separated by hωc and in each Landau level states with larger angular momentum are localized ¯ further from the origin. We recall from Eq.( 4) that wavefunctions with angular momentum m are localized [6] near a circle with radius Rm = 2(m + 1)ℓ. (Note that for large m the separation between adjacent values of Rm is ℓ2 /Rm << ℓ.) In the strong magnetic field limit the confinement potential does not mix different Landau levels. Since there is only one state with each angular momentum in each Landau level the only effect of the confinement potential is to increase the energy of the symmetric gauge eigenstates when Rm becomes larger than ∼ R. The typical situation is illustrated schematically in Fig. [5]. Here the n = 0 and n = 1 Landau levels are occupied in the bulk and the chemical potential µ lies in the gap ∆ = hωc between the highest energy occupied Landau level (E = 3¯ ωc /2) and the lowest ¯ h energy unoccupied Landau level (E = 5¯ ωc /2). In this section we are interested only in the h ground state and the low energy excited states obtained by making one or more particle-hole excitations at the edge. We will discuss only the simplest situation where a single Landau level crosses the chemical potential at the edge of the system and the analogous single branch situations in the case of the fractional quantum Hall effect [34]. We will also neglect the spin degree of freedom of the electrons, which figured so prominently in the previous section. An important property of the ground state of the non-interacting electron system in the case of interest, is that it remains an exact eigenstate of the system (but not necessarily the ground state!) when interactions are present. That is because the total angular momentum K for this state is N −1 M0 = m=0 m = N(N − 1)/2 15 (13) and all other states in the Hilbert space (truncated to the lowest Landau level) have larger angular momentum [35]. For large disks and total angular momentum near M0 the excitation energy of a non-interacting electron state will be ∆E = γM (14) where M ≡ K −M0 is the excess angular momentum and γ is the energy separation between single-particle states with adjacent angular momenta and energies near the Fermi energy. γ is related to the electric field, Eedge from the confining potential at the edge of the disk: γ = eEedge dRm = eEedge ℓ2 /R dm (15) This expression for γ can be understood in a more appealing way. In a strong magnetic field charged particles execute rapid cyclotron orbits centered on a point that slowly drifts in the direction perpendicular to both the magnetic field and the local electric field. For an electron at the edge of the disk the velocity of this ‘E cross B’ drift is vedge = cEedge /B. The energy level separation can therefore be written in the form γ = hvedge /Redge = h/T ¯ (16) where T is the period of the slow drift motion of edge electrons around the disk, in agreement with expectations based on semiclassical quantization. Since the excitation energy depends only on the angular momentum increase compared to the ground state it is useful to classify states by M. It is easy to count the number of distinct many-body states with a given value of M as illustrated in Fig. [6]. For M = 1 only one many-particle state is permitted by the Pauli exclusion principle; it is obtained by promoting the ground state electron with m = N − 1 to m = N. For M = 2, particle hole excitations are possible from m = N − 1 to m = N + 1 and from m = N − 2 to m = N. In general M many-particle states with excess angular momentum M can be created by making a single-particle hole excitation of the ground state. For M ≥ 4 additional states can be created by making multiple particle-hole excitations. The first of these is a state with two particle-hole excitations that occurs at M = 4 and is illustrated in Fig. [6]. 16 B. Many-Body Wavefunction Picture We now discuss the edge excitation spectrum of interacting electrons using a language of many-particle wavefunctions. For the case of the integer quantum Hall effect we will essentially recover the picture of the excitation spectrum obtained previously for non-interacting electrons by counting occupation numbers. We could have used the Hartree-Fock approximation and occupation number counting to generalize these results to interacting electrons. However, the Hartree-Fock approximation is completely at sea when it comes to the fractional case. Discussions of the fractional edge using an independent electron language can be comforting but can also be misleading. Nevertheless, we will see that there is a one-toone correspondence between the edge excitation spectrum for non-interacting electrons at integer filling factors and the fractional edge excitation spectrum. Many-electron wavefunctions where all electrons are confined to the lowest Landau level must be sums of products of one-particle wavefunctions from the lowest Landau level. From Eq. (4) it follows that any N electron wavefunction has the form Ψ[z] = P (z1 , . . . , zN ) ℓ exp (−|zℓ |2 /4), (17) where we have adopted ℓ as the unit of length and P (z1 , . . . , zN ) is a polynomial in the twodimensional complex coordinates. This property [36] of the wavefunctions will be exploited in this section. The first important observation is that since Ψ[z] is a wavefunction for many identical fermions it must change sign when any two particles are interchanged, and therefore must vanish as any two particles positions approach each other. Since P (z1 , . . . , zN ) is a polynomial in each complex coordinate it follows [37] that P (z1 , . . . , zN ) = i 0) electrons, nL (x) and nR (x): E[nL , nR ] = E0 + dx αLL 2 αRR 2 δnL (x) + δnR (x) + αLR δnL (x) δnR (x) . 2 2 (25) It is, perhaps, not completely obvious that the density provides a complete parameterization of the low-energy excitations, and indeed in the fractional Hall case there are situations where the analog of Eq. (25) is incorrect. [45] Here αLL , αLR and αRR are determined by the second derivatives of the energy per unit length with respect to nL and nR for a uniform system and can be determined in principle by a microscopic calculation. δnL (x) and δnR (x) are differences of the density from the ground state density. Note that we have as a convenience chosen the chemical potential to be zero in dropping a term proportional to dx(δnL (x) + δnR (x)). We start by considering the case where αLR = 0 so that the left-moving electrons and right moving electrons are decoupled. Focus for this case on the energy of the right moving electrons. We Fourier expand the density and note that 21 dx δn2 (x) = R 1 n−qR .nqR L q=0 (26) so that the energy can be written in the form ER = E0 + αLL 2L n−qR nqR . (27) q=0 The energy above can be used as an effective Hamiltonian for low-energy long-wavelength excitations. The simplification at the heart of the Luttinger liquid theory is the observation that when the Hilbert space is truncated to include only low-energy, long-wavelength excitations (in particular when the number of left-moving and right-moving electrons is fixed) Fourier components of the charge density do not commute. For example consider the second quantization expression for nqR in terms of creation and annihilation operators with k > 0: c† ck . k+q nqR = (28) k>0 An example of the dependence of the effect of products of these operators on the order in which they act is more instructive than the actual algebraic calculation of the commutators. Note for example that n−qR |Ψ0 = 0 (29) where q > 0 and |Ψ0 is the state with all right-going electron states with k < kF occupied and all right-going states with k > kF empty. (The alert reader will have noticed that this state of ‘right-going’ electrons corresponds precisely to the ‘maximum density droplet’ states that occur in the quantum Hall effect.) n−qR annihilates this state because there are no right-electron states with a smaller total momentum than |Ψ0 . On the other hand for q = M2π/L, nqR |Ψ0 yields a sum of M terms in which single-particle hole excitations have been formed in |Ψ0 . For example, if we represent occupied states by solid circles and unoccupied states by open circles, as in Fig. (6), for M = 2 we have nqR |Ψ0 = | . . . • • ◦ •| • ◦ ◦ . . . +| . . . • • • ◦| ◦ • ◦ . . . . 22 (30) Each of the M terms produced by nqR |Ψ0 is mapped back to |Ψ0 by n−qR . Therefore nqR n−qR |Ψ0 = 0 whereas n−qR nqR |Ψ0 = M|Ψ0 . The general form of the commutation relation is readily established by a little careful algebra [44]: [n−q′ R , nqR ] = qL δq,q′ . 2π (31) This holds as long as we truncate the Hilbert space to states with a fixed number of rightgoing electrons and assume that states far from the Fermi edge are always occupied. We can define creation and annihilation operators for density wave excitations of rightgoing electrons. For q > 0 aq = 2π n−qR qL (32) a† = q 2π nqR qL (33) With these definitions Eq. (31) yields [aq′ , a† ] = δq,q′ q (34) so that the density waves satisfy bosonic commutation relations. Also note that qL ˆ [M, aq ] = − aq 2π qL † ˆ a [M, a† ] = q 2π q (35) (36) ˆ where M is the total angular momentum operator. The contribution to the Hamiltonian from right-going electrons is therefore h vq a† aq ¯ q HR = (37) q>0 where v= αRR 1 d2 E0 1 dµR = = 2 2π¯ h 2πL¯ dnR h 2π¯ dnR h (38) At low-energies the system is equivalent to a system of one-dimensional phonons traveling to the right with velocity v. In the limit of non-interacting electrons 23 v= h kF ¯ ≡ vF m∗ (39) as expected. Without interactions between left and right-moving electrons a Luttinger liquid is quite trivial. In particular the ground state (|Ψ0 ) is a single-Slater determinant with a sharp Fermi edge. For one-dimensional electron gas systems the interesting physics [44] occurs only when left and right-moving electrons are allowed to interact. Most notably, arbitrarily weak interactions destroy the sharp Fermi edge that is the hallmark of Fermi liquids and that survives interactions in higher dimensions. In the case of quantum Hall edges, however, the above restriction to electrons moving in only one direction is not a temporary pedagogical device. The model with only right moving electrons discussed above can be taken over mutatis mutandis as a model of the edge excitations for an electron system with ν = 1. The role played by the one-dimensional electron density is taken over by the integral of the twodimensional electron density along a line perpendicular to the edge. In this way we arrive at the same bosonized picture of the ground state and low-lying excitations at the edge of a quantum Hall system as we reached previously by arguing in terms of many-particle wavefunctions. The single boson states which appeared there are replaced by the states of the chiral phonon system which has modes with only one sign of momentum and velocity. For ν = 1 the analysis applies whether or not the electrons interact. We now turn our attention to a discussion of the fractional case. Do all steps of the above discussion generalize? We can argue that if we are interested only in low-energy long-wavelength excitations, the energy can be expressed in the form E = E0 + α n−q nq . 2L q=0 (40) As we comment later, this expression can fail at the edge of fractional quantum Hall systems although it is appropriate for ν = 1/m. What about the commutator? There is an important difference in the line of argument in this case, since single-particle states far from the edge of the system are not certain to be occupied. Instead the average occupation number is 24 ν = 1/m and there are large quantum fluctuations in the local configuration of the system even in the interior. However, we know [39] from the discussion in terms of many-body wavefunctions in the previous section that the low-energy excitations at ν = 1/m can be described as the excitations of a boson system, exactly like those at ν = 1, which suggests that something like Eq. (31) must still be satisfied when the Hilbert space is projected to low energies. If we replace the commutator by its expectation value in the ground state we obtain [n−q′ , nq ] = ν · qL δq,q′ 2π (41) which differs from Eq. (31) only through the factor ν. It seems clear for the case of ν = 1/m this replacement can be justified on the grounds that the interior is essentially frozen (but in this case not simply by the Pauli exclusion principle) at excitation energies smaller than the gap for bulk excitations. What we need to show is that Eq. (41) applies as an operator identity in the entire low-energy portion of the Hilbert space. Below, however, we follow a different line of argument. Appealing to the microscopic analysis in terms of many-body wavefunctions we know that the excitation spectrum for ν = 1/m is equivalent to that of a system of bosons. We conjecture that the commutator [n−q′ , nq ] =∝ qδq,q′ . To determine the constant of proportionality we will require that the rate of change of the equilibrium edge current with chemical potential be eν/h. From the edge state picture of the quantum Hall effect discussed in Section II, it is clear that this is equivalent to requiring the Hall conductivity to be quantized at νe2 /h. Since our theory will yield a set of phonon modes that travel with a common velocity v it is clear that the change in equilibrium edge current is related to the change in equilibrium density by δI = evδn. (42) When the chemical potential for the single edge system is shifted slightly from its reference value (which we chose to be zero) the grand potential is given by 25 E[n] = E0 + µδn + α (δn)2 2 (43) Minimizing with respect to δn we find that δµ α (44) ev δI = δµ α (45) δn = so that In order for this to be consistent with the quantum Hall effect (δI = (eν/h)δµ) our theory must yield a edge phonon velocity given by v= α · ν. h (46) The extra factor of ν appearing in this equation compared to Eq. (38) requires the same factor of ν to appear in Eq. (41). We discuss below the qualitative changes in the physics [46,43] of fractional edge states which are implied by this outwardly innocent numerical factor. It is worth remarking that the line of argument leading to this specific chiral Luttinger liquid theory of the fractional quantum Hall effect is not completely rigorous. In fact we know that this simplest possible theory with a single branch of chiral bosons does not apply for all filling factors [39,46,47], even though (nearly) all steps in the argument are superficially completely general. The reader is encouraged to think seriously about what could go wrong with our arguments. Certainly the possibility of adiabatically connecting all low-energy states with corresponding states of the non-interacting electron system, available for one-dimensional electron gases and for quantum Hall systems at integer filling factors but not at fractional filling factors, adds confidence when it is available. In our view, the microscopic many-particle wavefunction approach that establishes a one-to-one mapping between integer and fractional edge excitations (for ν = 1/m!) is an important part of the theoretical underpinning of the Luttinger liquid model of fractional Hall edges. Once we 26 know that the edge excitations map to those of a chiral boson gas and that the fractional quantum Hall effect occurs, it appears that no freedom is left in the construction of a lowenergy long-wavelength effective theory. An important aspect of Luttinger liquid theory is the expression for electron field operators in terms of bosons [44]. This relationship is established by requiring the exact identity ˆ ˆ [ρ(x), ψ † (x′ )] = δ(x − x′ ) ψ † (x′ ) (47) to be reproduced by the effective low-energy theory. This equation simply requires the electron charge density to increase by the required amount when an electron is added to the system. The electron creation operator should also be consistent with Fermi statistics for the electrons: ψ † (x), ψ † (x′ ) = 0. (48) In order to satisfy Eq. (47), the field operator must be given by −1 ˆ ψ † (x) = ceiν φ(x) (49) where dφ(x)/dx = n(x) and c is a constant that cannot be determined by the theory. The factor of ν −1 in the argument of the exponential of Eq. (49) is required because of the factor of ν in the commutator of density Fourier components that in turn was required to make the theory consistent with the fractional quantum Hall effect. When the exponential is expanded the k − th order terms generate states with total boson occupation number k and are multiplied in the fractional case by the factor ν −k ; multi-phonon terms are increased in relative importance. It is worth remarking [46] that the anticommutation relation between fermion creation operators in the effective theory is satisfied only when ν −1 is an odd integer. This provides an indication, independent of microscopic considerations, that the simplest single-branch chiral boson effective Hamiltonian can be correct only when ν = 1/m for odd m. Wen [43] has surveyed, using this criterion, the multi-branch generalizations of the simplest effective Hamiltonian theory which are possible at any given rational filling factor. 27 His conclusions are consistent with arguments [39] based on the microscopic theory of the fractional quantum Hall effect. Eq. (49) has been carefully checked numerically [48] and appears to be correct. The ν −1 factor leads to predictions of qualitative changes in a number of properties of fractional edges. The quantity that is most directly altered is the tunneling density-of-states. Consider, for example, the state created when an electron, localized on a magnetic length scale, is added to the ground state at the edge of a N− electron system with ν = 1/m: a† √ n |ψ0 nν n>0 1 phonon term 2 phonon terms =1+ + + .... ν 1/2 ν ˆ ψ † (0)|Ψ0 ∼ exp − (50) The tunneling density states is given by a sum over the ground and excited states of the N + 1 particle system: A(ǫ) = n δ(En − E0 − ǫ)| Ψn |ψ † (0)|Ψ0 |2 (51) Because of the increased weighting of multiphonon states, which become more numerous at energies farther from the chemical potential, the spectral function is larger at larger ǫ − µ in the fractional case. An explicit calculation [46,43] yields a spectral function that grows like (ǫ − µ)ν −1 −1 . It is intuitively clear that the spectral function should be small at low-energies in the fractional case since the added electron will not share the very specific correlations common to all the low-energy states. It is amazing that by simply requiring the low-energy theory to be consistent with the fractional quantum Hall effect we get a very specific prediction for the way in which this qualitative notion is manifested in the tunneling density of states. V. ACKNOWLEDGMENTS The ideas discussed here have been shaped by discussions with members of the condensed matter theory group at Indiana University, especially M. Abolfath, S.M. Girvin, C. Hanna, 28 R. Haussmann, S. Mitra, K. Moon, J.J. Palacios, D. Pfannkuche, E. Sorensen, K. Tevosyan, K. Yang, and U. Z¨ licke. Discussions with L. Brey, R. Cote, H. Fertig, M. Fisher, M. u Johnson, C. Kane, L. Martin, J. Oaknin, C. Tejedor, S.R.-E. Yang and X.-G. Wen are also gratefully acknowledged. The responsibility for surviving misapprehensions rests with me. This work was supported by the National Science Foundation under grant DMR-9416906. 29 REFERENCES [1] K. v. Klitzing, G. Dorda, and M. Pepper, Phys. Rev. Lett. 45, 494 (1980); D.C. Tsui, H.L. St¨rmer, and A.C. Gossard, Phys. Rev. Lett. 48, 1761 (1982). o [2] A thorough introduction to the quantum Hall Effect is available from several books on the topic: The Quantum Hall Effect, edited by Richard E. Prange S.M. Girvin (SpringerVerlag, New York, 1987) emphasizes major early developments in theory and experiment; Introduction to the Theory of the Integer Quantum Hall Effect, by M. Janssen, O. Viehweger, U. Fastenrath, and J. Hajdu (VCH Verlagsgesellschaft, Weinheim, 1994) emphasizes localization physics important for the integer quantum Hall effect; The Fractional Quantum Hall Effect by T. Chakraborty and P. Pietil¨inen emphasizes numerical a exact diagonalization studies of the fractional quantum Hall effect; Quantum Hall Effect: A Perspective (Kluwer, Boston, 1989) by A.H. MacDonald is a reprint volume emphasizing the scientific context in which the quantum Hall effect discovery occurred; The Quantum Hall Effect, edited by Michael Stone is a reprint volume emphasizing topological aspects of the quantum Hall effect. [3] Marcus Kasner and A.H. MacDonald, Physica B 212, 289 (1995); ibid preprint (1995). [4] N. Read and Subir Sachdev, Phys. Rev. Lett. 75, 3509 (1995). [5] Tin-Lun Ho, Phys. Rev. Lett. 73, 874 (1994). [6] A.H. MacDonald, in Les Houches, Session LXI, 1994, Physique Quantinque Mesoscopique, edited by E. Akkermans, G. Montambeaux, and J.L. Pichard (Elsevier, Amsterdam, 1995). [7] A.H. MacDonald in Proceedings of the 7th Brazilian Workshop on Semiconductor Physics, Brazilian J. Phys. 26, to appear (1996). [8] For related ‘edge-state’ explanations of the quantum Hall effect see R.B. Laughlin, Phys. Rev. B 23, 5632 (1981); B.I. Halperin, Phys. Rev. B 25, 2185 (1982); A.H. MacDonald 30 and P. Streda, Phys. Rev. B 29, 1616 (1984); M. Buttiker, Phys. Rev. B 38, 9375 (1988). [9] P. Stˇeda, J. Phys. C, 15, L717 (1982); L. Smrˇka and P. Stˇeda, J. Phys. C, 10, 2153 r c r (1977); A. Widom and T.D. Clark, J. Phys. D, 15, L181 (1982). [10] See for example Gerald Bastard, Wave Mechanics Applied to Semiconductor Het´ erostructures, (Les Editions de Physique, Paris, 1988). [11] See for example, A.H. MacDonald, Surface Science, 229 1 (1990) and work cited therein. [12] F.D.M. Haldane, Chapter 7 in The Quantum Hall Effect, edited by R.E. Prange and S.M. Girvin (Springer-Verlag, New York, 1987). [13] A.H. MacDonald and D.M. Murray, Phys. Rev. B 31, 2291 (1985). [14] R. B. Laughlin, Phys. Rev. Lett. 50, 1395 (1983). [15] π φm |r 2 |φm = 2π(m + 1). [16] A.H. MacDonald, H.A. Fertig, and Luis Brey, preprint (1995). [17] See for example Hal Tasaki, Phys. Rev. Lett. 73, 1158 (1994) and work cited therein. [18] S.L. Sondhi, A. Karlhede, S.A. Kivelson, and E.H. Rezayi, Phys. Rev. B 47, 16419 (1993). [19] R. Rajaraman, Solitons and Instantons (North-Holland, Amsterdam, 1989). [20] T.H.R. Skyrme, Proc. R. Soc. A262, 233 (1961); A.A. Belavin and A.M. Polyakov, JETP Lett. 22, 245 (1975). [21] See for example, Eduardo Fradkin, Field Theories in Condensed Matter Physics, (Addison-Wesley, Redwood City, 1991). [22] D.-H. Lee and C.L. Kane, Phys. Rev. Lett. 64, 1313 (1990). 31 [23] S.-C. Zhang, T.H. Hansson, and S.A. Kivelson, Phys. Rev. Lett. 62, 82 (1989). [24] N. Read, Phys. Rev. Lett. 62, 86 (1989). [25] K. Moon, H. Mori, Kun Yang, S.M. Girvin, and A.H. MacDonald, L. Zheng, D. Yoshioka, and Shou-cheng Zhang, Phys. Rev. B 51, 5138 (1995). [26] Kun Yang and A.H. MacDonald, Phys. Rev. B 51, 13475 (1995). [27] E.H. Rezayi, Phys. Rev. B 36, 5454 (1987); Phys. Rev. B 43, 5944 (1991). [28] R. Tycko, S. E. Barrett, G. Dabbagh, L. N. Pfeiffer, K. W. West, Science 268, 1460 (1995); S. E. Barrett, G. Dabbagh, L. N. Pfeiffer, K. W. West, R. Tycko, Phys. Rev. Lett. 74, 5112 (1995). [29] H.A. Fertig, L. Brey, R. Cˆt´, and A.H. MacDonald, Phys. Rev. B 50, 11018 (1994); oe ibid, submitted to Phys. Rev. B (1996). [30] A. Schmeller and J.P. Eisenstein, Phys. Rev. Lett. 75, 4290 (1995). [31] E.H. Aifer, B.B. Goldberg, and D.A. Broido, preprint (1995). [32] L. Brey, H.A. Fertig, R. Cote, and A.H. MacDonald, Phys. Rev. Lett. 75, 2562 (1995). [33] Lack of time and space limits the scope of our discussion to systems without internal boundaries that have a single outer edge. There is, appropriately, considerable interest in annular and equivalent geometries that have inner and outer edges carrying currents in opposite directions. These geometries allow the problem of scattering between edges by disorder, related to transport experiments, to be considered. For work on this problem see C.L. Kane and M.P.A. Fisher, Phys. Rev. Lett. 68, 1220 (1992); Phys. Rev. B 46, 15233 (1992); M. Ogata and H. Fukuyama, Phys. Rev. Lett. 47, 4631 (1993); K.A. Matveev, D. Yue, and L.I. Glazman, Phys. Rev. Lett. 49, 1966 (1994); K.A. Matveev, D. Yue, and L.I. Glazman, Phys. Rev. B 49, 1966 (1994). [34] For recent work on ‘multi-branch’ systems see C.L. Kane and Matthew P.A. Fisher, 32 Phys. Rev. B 51, 13449 (1995) and work cited therein. [35] For an elementary discussion of some properties of this maximum density droplet state see A.H. MacDonald, S.R. Eric Yang, and M.D. Johnson, Aust. J. Phys. 46, 345 (1993). [36] The Hilbert space of the lowest Landau level is the Hilbert space of analytic functions: S.M. Girvin and T. Jach, Phys. Rev. B 29, 5617 (1984). [37] A.H. MacDonald and D.B. Murray, Phys. Rev. B 32 2291 (1985). [38] Michael Stone, Phys. Rev. B 42, 8399 (1990); Michael Stone, Ann. Phys. (NY) 207, 38 (1991); Michael Stone H.W. Wyld, and R.L. Schult, Phys. Rev. B 45, 14156 (1992). [39] For a more general discussion see A.H. MacDonald, Phys. Rev. Lett. 64, 220 (1990). [40] The antisymmetry requirement generally requires P [z] must be an odd function of zi −zj for any i and j so that P [z] contains only odd terms in a Taylor series expansion of its dependence on this difference coordinate. The coefficient of the (zi − zj )1 term must vanish if the pair never has relative angular momentum equal to one so that the leading term is (zi − zj )3 . It then follows from analyticity that each difference coordinate to the third power is a factor of P [z]. [41] A.H. MacDonald and M.D. Johnson, Phys. Rev. Lett. 70, 3107. [42] See for example, George E. Andrews, The Theory of Partitions (Addison-Wesley, Reading, 1976). A famous theorem by Hardy and Ramanujan in the theory of partitions gives an exact formula expressing g(M) in terms of a finite sum. [43] For reviews see X.G. Wen, Int. J. Mod. Phys. B6, 1711 (1992); X.G. Wen, to appear in Advances in Physics (1995). [44] See for example, J. S´lym, Adv. Phys. 28, 201 (1979); F.D.M. Haldane, J. Phys. C 14, o 2585 (1981); G.D. Mahan, Many-Particle Physics, (Plenum, New York, 1990) Chapter 4; J. Voit, preprint (1995). [To appear in Reports on Progress in Physics.] K. Sch¨nhammer o 33 and V. Meden, submitted to Am.J. Phys. (1995). [45] The microscopic analysis of fractional quantum Hall edges in Ref. [39] shows that the parameterization of the energy density at the edge in terms of the linear charge density is successful only for the simple case where ν = 1/m that we have discussed in these notes. [46] X.G. Wen, Phys. Rev. B 41, 12838 (1990); D.H. Lee and X.G. Wen, Phys. Rev. Lett. 66, 1765 (1991); X.G. Wen, Phys. Rev. B 44, 5708 (1991). [47] M.D. Johnson and A.H. MacDonald, Phys. Rev. Lett. 67, 2060 (1991). [48] J.J. Palacios and A.H. MacDonald, Phys. Rev. Lett. 76, 118 (1996) and work cited therein. 34 FIGURES FIG. 1. A large but finite two-dimensional electron gas. In panel (a) the chemical potential lies in a gap and the only low-energy excitations are localized at the edge of the system. In panel (b) the chemical potential lies in a mobility gap so that there are low-energy excitations in the bulk but they are localized away from the edge. In panel (c) a net current is carried from source to drain by having local equilibria at different chemical potentials on upper and lower edges. FIG. 2. Illustration of a Skyrmion spin texture. At the center of the Skyrmion m points in the ˆ down (−ˆ) direction. Far from the center of a Skyrmion m points in the up (ˆ) direction. Along z ˆ z a ray at angle θ in a circular coordinate system defined with respect to the Skyrmion center, m ˆ rotates about an axis in the (sin(θ) − cos(θ) direction from −ˆ to z . At fixed r the x − y projection z ˆ ˆ ˆ of m has fixed magnitude and rotates by ±2π when the angular coordinate winds by ±2π. At ˆ r = λ, m lies entirely in the x − y plane. ˆ ˆ ˆ FIG. 3. Knight shift measurements by Barrett et al. of the spin polarization of a two-dimensional electron gas near filling factor ν = 1. Here S = A = K + 1/2 so that the experiment is consistent with K = 3 for this sample. The dashed line in this figure shows the dependence of spin-polarization on filling factor expected for non-interacting electrons and, in the Hartree-Fock approximation, also for interacting electrons. The spin-polarization is assumed to be proportional to the Knight shift of the 71 Ga nuclear resonance and to be complete at ν = 1. (After Ref. [28] FIG. 4. Dependence of spin-polarization P on filling factor for Skyrme lattice states. Here g is the ratio of the Zeeman energy to the characteristic interaction energy e2 /ℓ and the values chosen are typical of experimental systems. The open and closed circles are experimental results of Barrett et al.. The legends indicate the nature of the Skyrme lattice state: the SLA state is a square lattice state with opposing Skyrmion orientations; the TLF state is a triangular lattice state with aligned Skyrmion orientations. (After Brey et al. in Ref. [32]) 35 FIG. 5. Schematic spectrum for non-interacting electrons confined to a circular disk in a strong magnetic field. In the limit of large disks the dependence of the energy on m can usually be considered to be continuous. The situation depicted has Landau level filling factor ν = 2 in the bulk of the system. The low-energy excitations are particle-hole excitations at the edge of the system. FIG. 6. Non-interacting many electron eigenstates for small excess angular momentum M specified by occupation numbers for the single-particle states with energies near the chemical potential µ. The vertical bars separate single-particle states with ǫm < µ from those with ǫm > µ. A solid circle indicates that nm = 1 in both the ground state and in the particular excited state; a shaded circle indicates that nm = 1 in the particular excited state but not in the ground state; an empty circle indicates that nm = 0. 36 TABLES TABLE I. Quantum occupation numbers in boson and fermion descriptions for edge excitations with small excess angular momentum M . gM is the number of states with excess angular momentum M . The fermion occupation numbers are relative to the maximum density droplet state. Only non-zero values are listed for both fermion and boson descriptions. L = N − 1 is the highest angular momentum that is occupied in the maximum density droplet state. M gM Fermion Description Boson Description 1 1 nL+1 = 1, nL = −1 n1 = 1 2 2 nL+2 = 1, nL = −1; nL+1 = 1, nL−1 = −1 n2 = 1; n1 = 2 3 3 nL+3 = 1, nL = −1; nL+2 = 1, nL−1 = −1 n3 = 1; n2 = 1, n1 = 1; nL+1 = 1nL−2 = −1 n1 = 3 nL+4 = 1, nL = −1; nL+3 = 1, nL−1 = −1 n4 = 1; n3 = 1, n1 = 1; n2 = 2 nL+2 = 1, nL−2 = −1; nL+1 = 1, nL−3 = −1 n2 = 1, n1 = 2; n1 = 4 4 5 nL+2 = 1, nL+1 = 1, nL = −1, nL−1 = −1 37 20 15 (S=A=3.4+/-0.3) 10 71Ga Well Shift (kHz) Interacting Electrons 5 T=1.55 K, H=7.05 Tesla Independent Electrons (S=A=1) 0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 ν ... . . . gM=1 = 1 • M=2 ... ... ... . . . gM=2 = 2 ... ... • M=1 • M=4 . . . gM=4 = 5 ... ... EDGE REGION R (a) dA dl EQUILIBRIUM µ IN GAP EDGE REGION (b) EQUILIBRIUM µ IN MOBILITY GAP LOCALIZED STATES µ1 µ2 (c) LOCAL EQUILIBRIUM ON ISOLATED EDGES I I µ3 µ4 n=2 n=1 n=0 empty ∆ occupied occupied µ 1.0 SLA, g*= 0.015 TLF, g*= 0.015 SLA, g*= 0.020 Indep. electrons Experiment Experiment Spin Polarization, P 0.8 0.6 0.4 0.2 0.0 0.6 0.8 1.0 1.2 ν 1.4 1.6 1.8