This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
= 27r^(q,coMq)pqxt
(5.30)
We may relate this to the dielectric function as follows (Nozieres and Pines 1958). Maxwell's equations in vacuum give for the electric-field Fourier component Eq e o q.E q = e(p- t + < p q » (5-31) In the dielectric description of the continuous medium, the corresponding Maxwell equation gives the response to the external charge density: 80q-e{q9Q>)Eq = ep*«
(5.32)
Comparison of (5.31) and (5.32), with use of (5.28), establishes the relation 6(q, co) = [1 + 2nv(q)3f(q9 co)] ~ *
(5.33)
Substitution of (5.27) gives the useful alternative expression e(q, co)=l-
27n;(q)^0(q, co)
(5.34) 169
Electronic surface states and dielectric functions Equation (5.34) is the general RPA expression for the dielectric function of the 3D electron gas. It could be discussed at considerable length, but we restrict attention to the form that applies at relatively long wavelengths and low temperatures, q « qF and T«TF. Here qF is the Fermi wave vector and TF the Fermi temperature, given by kBTF = eF = h2qF/2m. These requirements are not too restrictive in a metal, for which qF ~ I/a, where a is the interatomic spacing, and TF >i 104 K. The appropriate limiting form of @° is given in Problem 5.5, and is repeated here for convenience: 2n<3°(q9 co) ~ nVq2/mco2
(5.35)
where n is the electron density. Equations (5.11), (5.34) and (5.35) now give s(q,co)=l-co2p/co2
(5.36)
col = ne2/som
(5.37)
with which will be recognised as the plasma frequency of the electron gas. Equation (5.36) shows that a(q, co) is negative for co < cop and positive for co > cop. As (5.33) shows, the zero of e(q, co) at co = cop corresponds to a pole of the Green function ®(q, co), which confirms that cop is an excitation frequency of the system. What has been done is hardly the simplest way of deriving (5.36), and a more straightforward method is given in §5.3.3. However, the results are fundamental, and our derivation may be used to bring out the nature of the approximations made in simpler approaches. It was assumed that the single-electron eigenfunctions were plane waves, so that strictly speaking the results apply for a bulk electron gas. It may be noted that in principle the expressions derived exhibit spatial dispersion, that is, ^-dependence of the dielectric function; it is only in the long-wavelength limit that (5.36) applies and spatial dispersion is absent. Equation (5.36) represents only the lowest-order term of a Taylor expansion in q/qF, the next term being of order (q/qF)2. In a coordinate-space formulation, this corresponds to the appearance of terms in V2Sn, where Sn is the deviation from equilibrium density. This term will be seen to arise naturally in the hydrodynamic description to be given in §5.3.3. In the long-wavelength limit on which (5.36) is based, Im[^°(q, co)] vanishes, as mentioned at the end of Problem 5.5. Thus in this lowest order the dielectric function is purely real and damping is absent. A simple way of introducing damping will be explained in §5.3.3. The functional dependence of (5.36) upon co is illustrated in Fig. 5.6. It may be noted for future reference that (5.36) applies to the electron or hole gas in a semiconductor such as InSb as well as to metals. For a semiconductor, cop depends on the doping since that determines the carrier density n. 170
Collective properties of the electron gas Typically the plasma frequency of a semiconductor is in the infrared, whereas the plasma frequency of a good metal is in the near ultraviolet. For example, for electrons in InSb with a density 10 23 m~ 3 , cop/2nc = 770 cm" 1 or fia)p = 96 meV, while ha>p = 6.2 eV for Na. The account given in §5.3.1 and in §5.3.2 up to this point applies to the infinite 3D electron gas, since plane-wave eigenfunctions are used. It is fairly easy to see in a qualitative way how the discussion is modified for the case when one or more boundaries are present, although the details become quite involved. Accounts are given by Dasgupta and Beck (1982), Flores and Garcia-Moliner (1984) and Del Sole (1986). For semi-infinite jellium, for example, the Hamiltonian is written in terms of electronic eigenfunctions of the semi-infinite medium rather than the plane waves appropriate for bulk jellium. The sums in (5.8) and in the subsequent derivation of the RPA dielectric function are then sums over the quantum numbers of these eigenfunctions. If the derivation is taken as far as the q2 terms that lead to V2dn terms in the bulk, the analysis yields boundary Fig. 5.6 Free-electron dielectric function of (5.36). The plasma frequency cop is given by (5.37) with the free-electron density of Ag. 0 10
20 ho) (eV)
-10-
-20-
171
Electronic surface states and dielectric functions
conditions on Sn. The need for these additional boundary conditions, usually referred to as ABCs, will be seen from a different point of view in §5.3.3. The use of ABCs was also a feature of the discussion of dipole-exchange modes of §4.4. In general, the calculations for surfaces yield a correction to the bulk value of the dielectric function that is significant only for the first few atomic layers. This leads to corrections to electromagnetic quantities, such as reflection coefficients, that are only a few percent of the values found from the Fresnel expressions with the bulk dielectric function. However, these corrections are large enough to be detected experimentally. As an example, Fig. 5.7 shows the calculated differential reflectivity for two polarisations from the surface of Si(lll)-2 x 1 (the notation is explained in §1.2). The wavefunctions used were based on the tight-binding approximation. Fig. 5.7 may be compared with the experimental results shown in Fig. 5.8. It is seen that in the frequency range /ko<0.8eV covered by the experiments the agreement is good, particularly as concerns the large anisotropy between the two different polarisations. There are some circumstances in which a thin charge sheet, which may be approximated as a 2D electron gas can be produced. A Lindhard type function may be used for the 2D electron gas, and a brief account is given in §5.3.4. The main use that will be made subsequently of (5.36) is in Chapter 6, where the properties of plasmon-polaritons will be discussed, and in Fig. 5.7 Calculated differential reflectivity {R - Rs)/Rs, where R s is the reflectivity of the 2 x 1 H covered surface, for Si(ll 1)—2 x 1. The light is normally incident, and two directions of polarisation are shown: full line, electric field parallel to [011]; broken line, electric field parallel to [211] [after Del Sole 1986]. 4.0
k
3.0 2.0 1.0 0 -1.0 -2.0
172
/is 1.0
A *^\ 2.0
** * * 3.0
^
4.0
v £(eV)
Collective properties of the electron gas
Chapter 7 for the analogous modes in superlattices. These are electromagnetically coupled modes of wavelength X comparable with that of visible light. For the bulk plasmon-polariton, therefore, the inequality q/qF« 1 applies and it is legitimate to use the lowest-order Taylor expansion that leads to (5.36). The simplest form of surface plasmonpolariton arises at the interface between a metal and vacuum, and is found for a range of frequencies below cop in which s(co) is negative. It is localised at the interface in the sense that its amplitude decays exponentially with distance away from the interface. The decay length is of the same order of magnitude as the wavelength X, so that the mode penetrates at least some hundreds of interatomic spacings into the metal. The modifications of £(co) from its bulk form due to the presence of the surface are of much shorter range than this. It is therefore usually sufficient to characterise Fig. 5.8 Experimental results for differential reflectivity of Si(lll)—2 x 1. The polarisations are the same as in Fig. 5.8 [after Chiraradia et al. 1984].
5.0-
4.0-
3.0
2.0-
1.0-
0.0
0.3
0.4
0.5 Energy (eV)
—i—
0.6
173
Electronic surface states and dielectric functions the metal as having the dielectric function (5.36) right up to the surface. It is this description that will form the basis of the discussion in Chapters 6 and 7. 5.3.3 Hydrodynamic description We give a very brief introducton to this topic; a full account is given by Boardman (1982). The electron gas is treated as a continuous fluid with density n(r, t) and velocity v(r, t)y with associated electric current density j(r, t)\ }=-ne\
(5.38)
where dependence on r and t is understood. The particle current density is HV, so n and v are related by the continuity equation dn/dt + V-(n\) = 0
(5.39)
The velocity is assumed to obey the acceleration equation m(
+ v . V )v = -e(E + v x B) - mv/r - (mP2/n)Wn \dt J
(5.40)
As in all hydrodynamics, the total derivative D/Dt = d/dt + vS7 (5.41) is used; the term v»Vv is nonlinear, and is usually negligible. The term — mv/r introduces the collision, or damping, time t. The final term results from the compressibility of the electron gas. The value of the parameter P is discussed by Boardman (1982). The compressibility term ultimately leads to a term in q2 in the dielectric function e(q, a>), and comparison with the q2 term found by expansion of (5.34) shows that p2 = 3i;pa>p/5co2, where vF = hq¥ /m is the Fermi velocity. In practice, the simpler form P2 = 3fp/5 is often assumed. It is a major advantage of the hydrodynamic method that the magnetic-field and damping terms can be included so readily. The dielectric function describes the linear response of the electron gas, and we therefore proceed by linearising (5.38) to (5.40). We treat the electric field as the driving force, so that E and v are first-order quantities, whereas n is written as a sum of zero- and first-order terms: n = n0 + n x (5.42) We consider a single-frequency field E exp( —icot), so d/dt is replaced by — ico. The linearised forms of (5.38) to (5.40) are then }=-noe\
(5.43)
-koHi-h noV-v = 0
(5.44) 2
icom\ = mv/r + eE + e\ xB + (mp /n0)Vn1 174
(5.45)
Collective properties of the electron gas Substitution of nx from (5.44) into (5.45) gives the basic equation for v: icom\ = mv/t + eE + e v x B - (im^2/co)V V • v = 0
(5.46)
We restrict attention for the moment to a bulk electron gas, for which it is sufficient to take E in the plane-wave form E exp(iq-r — icot). We also set the magnetic field equal to zero. The operator V is then replaced by iq, and (5.46) can be solved for v: v = (eE/m)(ico - 1/T - ij8 VA»)~ *
(5.47)
Equations (5.43) and (5.47) are what is needed to determine the dielectric function e(q, co). The relevant Maxwell equation is written in the alternative forms iqxH=j-ico£oE
(5.48)
and iq x H = - icoe0e(q, co)E (5.49) where H is the first-order magnetic field generated by the electron current j . In (5.48), j is substituted from (5.43) and (5.47); comparison of (5.48) and (5.49) then yields e(q, co) = 1 - colico2 + ico/i -
fi2q2)'l
( 5 - 50 )
where the plasma frequency cop was defined in (5.37). Equation (5.50) is the previous (5.36), generalised to include damping through T and spatial dispersion through the compressibility term P2q2. Although the method does not have the generality of the microscopic theory which was the subject of §§5.3.1 and 5.3.2 it has the characteristic advantages of a hydrodynamic approach that it is relatively simple and damping is easily included in a phenomenological way. The main restriction is that it is a long-wavelength method, valid for qa«\, where a is the interatomic spacing. At the end of §5.3.2, it was pointed out that the calculations presented there were specifically for a bulk material, since the electron states were labelled by wavevector q. In the present approach, (5.46) could be used for a finite specimen such as a semi-infinite medium or a film. However, the solution given in (5.47) holds only for the bulk medium, since it involves q. For a finite specimen, the presence of the term in VV- v means that (5.46) can be solved only if a boundary condition on v is applied at the surface or surfaces. For example, a description of the reflection of light from the surface of semi-infinite jellium involves solving Maxwell's equations together with (5.43) and (5.46) for the medium. The boundary conditions are continuity of tangential E and H together with the boundary condition on v. It is because the latter is additional to the electromagnetic conditions on E and H that it is referred to as an ABC. The question of 175
Electronic surface states and dielectric functions ABCs for the electron gas will not be discussed further here; the reader may refer to the reviews cited in §5.3.2. As mentioned in §5.3.2, for the applications to plasmon-polaritons to be made in Chapters 6 and 7, it is sufficient to describe the medium as having the dielectric function (5.36) right up to the surface. Since this dielectric function is free from spatial dispersion, the question of ABCs will not arise. However, for some applications, such as the calculation of attenuated total reflection (ATR), it is essential to retain the damping term. We therefore for later reference quote explicitly the form of (5.50) with damping but without spatial dispersion, namely 8(o)) = 1 - col(o)2 + ico/r)-1
(5.51)
The hydrodynamic method can be used to give a simple treatment of the effect of an applied static magnetic field B o . The term e\ x B o is retained in the equation of motion (5.46). A straightforward extension of the derivation we have outlined shows that the dielectric tensor becomes anisotropic, and takes the form £ = | ie2
ex
0 |
(5.52)
+ lM
(5.53)
where fn (m
(5.54) and £3 is given by (5.51). The z axis is taken in the magnetic field direction. The cyclotron frequency coc = eBJm has been introduced. In (5.52) damping is included but spatial dispersion has been neglected. It is seen from (5.52) that in the presence of the magnetic field B o the xy components of s take a gyrotropic form, similar to the magnetic susceptibility tensor of (4.15). 5.3.4 Two-dimensional electron gas In some circumstances it is possible to have an electron gas in the form of a charge sheet that is sufficiently thin that it can be treated as a 2D electron gas. Two important examples are a charge sheet at the surface of liquid helium and a charge sheet in a semiconductor. Liquid helium has a dielectric constant of about 1.06, which is sufficient for electrons to be trapped in the weak image potential at the surface. The areal density of the 2D electron gas trapped in this way can be varied by 176
Dielectric function of ionic crystals a static electric field applied normal to the surface. A review of the topic is given by Cole (1974). In semiconductors, a static electric field applied normal to the surface, as in a field-effect transistor, can attract a charge sheet of electrons or holes to the surface, and again the areal density can be varied by changing the field. In addition, electrons can be trapped at a heterojunction, like that between GaAs and A ^ G a ^ ^ A s for example. Further discussion of this form of trapping is deferred to Chapter 7. It was first pointed out by Stern (1967) that the RPA discussion of the 3D electron gas, which formed the subject of §§5.3.1 and 5.3.2, could be taken over with appropriate modifications for the 2D electron gas. In particular the Lindhard function @°(q\\, co) can be defined as a function of the 2D wavevector q^ and frequency co. It is given formally by the same expression as the 3D function, (5.22), but naturally the sum is over a 2D wave vector ky. The dielectric function continues to be given by (5.34). Stern gives the result of an explicit evaluation of @°(q\\, o). Fetter (1973) showed how the hydrodynamic approach, outlined in §5.3.3, can be applied for the 2D case. He included the compressibility term, analogous to the last term of (5.45), that leads to spatial dispersion. His paper includes calculations of various response functions of the 2D gas embedded in a 3D medium. Subsequently Karsono and Tilley (1977) generalised this work to the case where the embedding media are different on either side of the layer, and also discussed a line of mobile charge, i.e. the ID electron gas. In a later paper, Fetter (1974) discussed a periodic array of 2D electron-gas sheets; some results for a periodic array of charge lines are given by Karsono and Tilley (1977). Both Stern (1967) and Fetter (1973) discuss electromagnetically-coupled modes propagating parallel to the charge sheet and localised on the charge sheet. Stern exclusively, and Fetter mainly, treat the electromagnetic field within the electrostatic approximation, which is analogous to the magnetostatic approximation that formed the basis of Chapter 4. Rather than introduce the electrostatic approximation explicitly here, we defer further discussion to §6.2.4. There, retardation is included, so that the electromagnetic field is treated exactly. The results obtained by the electrostatic approximation are derived as a limiting case. 5.4 Dielectric function of ionic crystals The dielectric function of the electron gas, discussed in §5.3, will be central to the discussion of plasmon-polaritons in Chapter 6. The other main example that will feature there is the phonon-polariton, which arises from the coupling of light to the optic phonons of an ionic lattice. By way of preparation, we here derive the expression for the dielectric function 177
Electronic surface states and dielectric functions that will be needed. Our account is quite brief, and is based on the method introduced by Born and Huang (1985). Most solid-state textbooks deal with the topic, and Ashcroft and Mermin (1976) give a particularly full account. For simplicity, we restrict attention to an isotropic or cubic medium, in which all vector quantities, such as P and E, are parallel. We may then use the lattice dynamics of a ID diatomic lattice, given in §2.1.2, as our starting point, and what follows may be regarded as an extension to include coupling to the electric field. The discussion of anisotropic media does not involve any new points of principle, and is not given here. As in §2.1.2, we consider an infinite diatomic lattice in which masses m1 and m2 alternate. We now take the crystal to be ionic, so that the two sublattices are associated with opposite electric charges. The mode that couples to light is the long wavelength (q = 0) optic phonon at frequency coT, say. This follows from (2.16) for the relative displacement, which shows that the sublattices move in antiphase. Thus the optic phonon is associated with an oscillating dipole moment, and therefore couples to light. The relevant wavenumber is q ~ 2n/A, where k is the wavelength of the light; on the scale of the Brillouin-zone edge n/2a this is effectively q = 0. We denote by u the relative displacement of the two sublattices. The polarisation P contains a term proportional to u, as well as a term due to the electronic susceptibility rather than the ionic displacements: P = eo(au + xE)
(5.55)
As mentioned, for an isotropic medium P, u and E are parallel. In (5.55) E is the macroscopic electric field, the value found by averaging the local field Eloc over many unit cells. Evaluation of the constants of proportionality a and x depends upon details of the lattice dynamics and electronic polarisability, and need not concern us. The equation of motion for u is {-co2 - icoT)u = -cofu + j?Eloc
(5.56)
On the left-hand side a damping term with damping constant T is included; it arises from an assumed damping force — Tdu/dt. The first term on the right-hand side ensures that in the absence of coupling to the electric field the mode frequency is c%, and the second is the driving force due to the local field. In order to derive an expression for the dielectric function from (5.55) and (5.56) it is necessary to find the relation of E loc to E. This is discussed by Ashcroft and Mermin (1976), for example. What Born and Huang pointed out, however, is that this relationship must be linear. Equation (5.56) can therefore be replaced by an equation involving E rather than Eloc: (-co2 - icoT)u = -to\u + yE 178
(5.57)
Dielectric function of ionic crystals Equations (5.55) and (5.57) are readily solved for P : (5.58) —
CO —
10)1
The defining equation for the dielectric function is D = a0E + P = eoe(a>)E
(5.59)
Combination of (5.58) and (5.59) gives
+ \
2
COj — COZ —
)
(5.60)
ICOiJ
where £oo = l + Z
(5-61)
co£-Q)f = a y / ( l + x )
(5.62)
and Equation (5.60) is the conventional form; it is written in terms of e^, the high-frequency (electronic) dielectric constant, and coL, the LO (longitudinal-optical) phonon frequency. The zero-frequency value of £ is !
(5-63)
which can be recognised as the Lyddane-Sachs-Teller (LST) relation. The frequency dependence of (5.60) is illustrated in Fig. 5.9. For typical dielectrics, coT lies in the far infrared, generally with wT/2nc in the region 100 to 1000 cm" 1 . It was mentioned in §5.3.2 that in semiconductors the plasma frequency lies in the infrared. With suitable doping, it can be made close in value to the TO frequency coT, so the dielectric function includes both plasma and optic-phonon contributions. The total dielectric function can be found by combination of the derivations we have given. For completeness, we record the form quoted by Palik et al. (1976) for the dielectric tensor of a doped polar semiconductor in the presence of a static magnetic field B o in the z direction. The tensor has the form of (5.52), with co2 — CD2 — icoF
co[co2 — (co 4- i / t ) 2 ]
e 2 in the form of (5.54), a n d
•"•-0 +*^-*£*i> Following the implications of this relatively complicated expression would take us too far afield, however, and it will not be used subsequently. 179
Electronic surface states and dielectric functions 5.5 Problems 5.1 Give a proof of (5.15). 5.2 Prove the following commutation results, used in the derivation of (5.20).
where the a+ and a operators are fermion creation and annihilation operators satisfying (5.10).
Fig. 5.9 (a) Real and (b) imaginary parts of dielectric function e{co) for GaAs, which is accurately described by (5.60). Numerical values are coj/2nc = 269.2cm" 1 , coL/2nc = 292.8 cm" 1 , T/2nc = 2.5 cm" 1 and e^ = 10.9 (Kim and Spitzer 1979). In (b) the sharpest curve is drawn for the quoted value of F; the two broader curves, drawn for comparison purposes, correspond to values of 5 x 2 . 5 cm" 1 and 25 x 2.5 cm" 1 .
250
-200 -
-400 -
180
260
270 X 280
290 300 310 (i)/2nc (cm" 1 )
320
Problems 5.3 By extending the results of Problem 5.2, evaluate the commutator 0 k A + q.
Z
l7 k
( >qi.aiflq2.»2flq2-k'.ff2flqi+k'.
Show that on application of the RPA decoupling the expression becomes
Z "( 5.4 The Lindhard function ^°(q, co) is defined in (5.22). With the sum converted to an integral by means of (1.19), it is
_iv_ r (2TC)
J
At zero temperature the Fermi occupation factor is fk = 6(qF-\k\) where 6 is the step function, see (A.2) of the Appendix, and qF is the
6000 -
5000 Im(e) 4000 -
3000 -
2000 -
1000 -
267
268
269
270
271
181
Electronic surface states and dielectric functions Fermi wavevector. For the free electron gas, q% = 3n2n, where n is the density. Fetter and Walecka (1971) evaluate a zero-temperature Green function n° which is closely related to £#°(q,co). In fact, ^ ° is a retarded function (response function), so that as in (A.26) the correct form in the complex plane involves making the replacement co -* co + irj in (5.22). By contrast, Fetter and Walecka have the so-called causal function, in which the replacement is co -> co + irj sgn(co). This means that the real parts are the same, Re(^°) = Re(7i°), but the imaginary parts are different. Following Fetter and Walecka, prove that at zero temperature (q,a>)]=
2V
-9
f
d3/c%F-|q|) 1
1 2
(q«k + ^7 )
hco-
where 0> denotes the principal-value integral. Hence show that h2
(2TT)3 {
2K L
2K
L
\K
\K
2)
2/ J
In
1 + (V/K + ; 1 - (V/K: -f \K
where K: = g/gF and v = mco/hq\. 5.5 If the above expression in Problem 5.4 is applied to a description of plasma oscillations, the limit with v fixed as K —• 0 is required. Show that this limit is Re[^ u (q, co)]
mqF h2
V
2K2
as K -• 0 4TT3 3V2
that is, Re[^ u (q, co)] ~
nV q2
as q -»• 0 27rmco2 Note that it can also be proved that Im[^°(q, co)] = 0 in this limit.
182
6 Surface polaritons
In dealing with magnetic surface modes in Chapters 3 and 4 we found it useful to classify modes, to some extent, by the range of the dominant restoring force. Thus Fig. 4.1 shows that for large enough wavevector |qiJ >> 108 m " 1 , the short-range exchange forces determine the nature of the surface modes. Over a wide range of intermediate wavevectors, approximately 104 m " 1 < |c|M| < 107 m " 1 the dipole-dipole force is dominant, and the important magnetostatic approximation holds. The properties of the magnetostatic surface modes are found by solving Maxwell's equations without retardation in the presence of the frequency-dependent tensorial magnetic susceptibility % given by (4.15) to (4.17). Finally, for long-wavelength modes, |q | < 10 3 m~ 1 , retardation cannot be neglected, and it is necessary to solve the full form of Maxwell's equations with susceptibility %. This is the polariton, or electromagnetic, region, discussion of which has been deferred to this chapter. Most of this chapter is concerned with electromagnetic, or polariton, modes arising from frequency-dispersion in the dielectric function rather than the magnetic susceptibility. The origin of dispersion was discussed in Chapter 5. Generally it results from the presence of a resonant mode carrying a dipole moment, like the optical phonon in a polar crystal. It is possible to define an electrostatic region, analogous to the magnetostatic, in which properties of bulk and surface modes are found by solving Maxwell's equations without retardation. The calculations are outlined in §6.2.2, but in practice the electrostatic modes are not very important, and most of the chapter deals with genuinely electromagnetic modes. Furthermore, spatial dispersion will not be included except in the discussion of bulk and surface exciton-polaritons. In §§6.3 to 6.3 we discuss bulk and surface dielectric polariton modes. Section 6.4 is an introduction to the important topic of nonlinear effects. 183
Surface polaritons The experimental technique that has been most comprehensively applied is attenuated total reflection, or ATR, and this forms the subject of §6.5. Section 6.6 describes what has been done to investigate surface polaritons by some other experimental methods. Finally, §6.7 deals wih surface magnon-polari tons. The guiding of electromagnetic waves at interfaces, which is the general theme of this chapter, is a matter of considerable importance in applications, and there is a substantial engineering literature on the subject. The interested reader will find more about applications in Marcuse (1982) and in Stegeman and Wallis (1986). 6.1 Bulk polaritons 6.1.1 Phonon and plasmon polaritons The propagation of light through a bulk medium characterised by a frequency-dependent dielectric function is governed by Maxwell's equations. For a plane wave with all fields proportional to exp(iq-r — icot) propagating through a bulk isotropic medium, the equation V • D = 0 gives £(co)q-E = 0
(6.1)
This has the solutions that either e(co) = 0
(6.2)
q.E = 0
(6.3)
or In the absence of damping, the first of these gives co = coL for an ionic medium, with dielectric function given by (5.60), and co = cop for the plasma with the dielectric function of (5.36). These are the longitudinal modes; they have no surface counterpart, and will therefore not detain us.f The second solution, given by (6.3), is the transversality condition (E transverse to q). For this solution, the V x E and V x H equations yield in the usual way q2 = e(co)a)2/c2
(6.4)
which is the propagation equation for the transverse mode. If the damping term in s(a>) is neglected then it is an easy matter to draw the dispersion curve corresponding to (6.4). For the phonon case, with e(co) given by (5.60), the asymptotic and limiting forms are
f It may be noted that in a detailed treatment, as given by Pines and Nozieres (1966) for example, it emerges that the dielectric function for longitudinal modes is different from that for transverse modes.
184
Bulk polaritons q2~e(0)co2/c2 2
q -> oo
CO«CO T
co-+coT
2
q <0
COT<0)
q2 ~z^to2lc2
CD-* co
(6.5)
These are illustrated in Fig. 6.1. The reader will note in particular that for coT < co < coL, q is pure imaginary, and therefore this frequency interval appears as a stop band in Fig. 6.1. Comparison with the results derived for lattice dynamics in Chapter 2 makes it plausible that if a surface mode exists it will be within this stop band. It will be seen in §6.2 that there is indeed such a mode, namely the surface polariton. The dispersion curve drawn in Fig. 6.1 may be seen as resulting from the crossing of the phonon line co = coT and the photon line q = s1/2co/c, where e is imagined to change slowly from s(0) to e^. These modes interact strongly, and the crossover is therefore eliminated with repulsion of the dispersion curves, as shown schematically in the insert to Fig. 6.1. Thus the dispersion curves of Fig. 6.1 describe a mode of mixed phonon-photon character. Many years ago, the word polariton was coined to describe such a mixed mode, so one may say that Fig. 6.1 depicts the bulk polariton Fig. 6.1 Dispersion curve for bulk phonon-polariton in GaAs; numerical values quoted for Fig. 5.9. Asymptotic lines q ~ [8(0)]1/2
1.5-
0.5
1.0
2.0
3.0 cq/a)r
4.0
5.0
6.0
185
Surface polaritons dispersion curve. Nowadays polariton is used for any mixed mode involving a photon. The plasma dielectric function of (5.51) is the toT -+ 0 limit of the phonon function. The dielectric function, drawn in Fig. 5.6 for y = 0, and the plasmon-polariton curve, drawn as Fig. 6.2, can clearly be derived from the corresponding phonon curves by letting coT -• 0. For a doped polar semiconductor, s(co) has two resonances, one plasma-like at co = 0, and one phonon-like at co = coT. Correspondingly there are two branches in the polariton dispersion curve rather than the one of Fig. 6.2. This is illustrated for n-InSb in Fig. 6.3. The dispersion curves of Figs. 6.1 to 6.3 were drawn for the case when damping is neglected in the dielectric functions. Equation (6.4) is then a relation between two real variables q and co, and the dispersion curve is a natural representation. However, if damping is included, (6.4) relates two variables q and co which in principle are complex. There is then not much sense in discussing (6.4) in isolation, since the conditions of a given experiment will determine what trajectory is followed in the space of the two complex variables q and co. For example, in inelastic light (Raman or Brillouin) scattering in a bulk medium, the experimental constraint is that q is real, since it is the wavenumber difference between the incident and scattered light. The real part of co then appears as a line position, and the imaginary part as a linewidth. Conversely, it is sometimes the
Fig. 6.2 Bulk plasmon-polariton dispersion curve. 4.0
186
T
Bulk polaritons
case in attenuated-total-reflection (ATR) experiments that co is constrained to be real. With damping present, then, discussion of the dispersion relation without reference to experimental conditions is a rather sterile exercise, which we shall not undertake. In this section we have given a very brief account of bulk polaritons in the simple cases of an isotropic medium whose dielectric response is dominated by one or two resonances. These bulk modes were the subject of very active investigation, particularly by Raman scattering, from the mid-1960s. A good account of that work is found in various papers in the Conference Proceedings edited by Burstein and De Martini (1974). A number of extensions and generalisations of calculations outlined in this section were necessary. The theory of bulk polaritons in an anisotropic crystal with multiple resonances is given by Merten in Burstein and De Martini (1974), and several articles describe related experimental results. As we have seen in earlier chapters, for a theory of light-scattering intensities it is necessary to go beyond the solution of the homogeneous equations of motion, which in this case are Maxwell's equations together with the appropriate dielectric functions. Several articles in Burstein and De Martini (1974) give accounts of light-scattering theory; one approach is to use the linear-response method, which is summarised in §1.3.2, and in the Appendix, where additional references are quoted.
Fig. 6.3 Bulk polariton dispersion curve for n-InSb with doping 10 23 m 3 . The dielectric function e(a>) has parameters (Palik et al. 1976): coj/lnc = 180cm" 1 ; coL/2nc= 192 cm" 1 ; e^ = 15.7, m*/me = 0.022, leading to cop/2nc= 161cm" 1 . /
/
/ /
0)/(OT
/
1.5-
/ / / / /
/ /
J
1.0-
0 7-
1
.—
1
cq/a)r
10
187
Surface polaritons 6.1.2 Exciton-polaritons and resonant Brillouin scattering As mentioned, the exciton-polariton is the most important case in which spatial dispersion of the dielectric function must be taken into account. In this section we give an introduction to bulk exciton-polaritons. The corresponding surface modes are discussed in §6.2.5. An exciton is an electron-hole pair, bound by the Coulomb interaction. Excitons in their own right are a major field of study; introductory accounts are given by Kittel (1986) and Smith (1978), for example, and Rashba and Sturge (1982) give a comprehensive account. The type of exciton of interest here arises in a direct-gap, polar semiconductor, such as GaAs or ZnSe. The exciton is formed of an electron from near the bottom of the conduction band and a hole from near the top of the valence band. Since the valence band usually comprises a heavy-hole band, a light-hole band and a band split off by the spin-orbit coupling, and the bands are not strictly isotropic, the full theory of the exciton is quite complicated. We restrict attention to the simplest model, in which an electron of effective mass me is bound to a hole of mass mh. This model contains the basic physics, and is adequate for an account of much of the experimental data on exciton-polaritons. The exciton wavefunction spreads over many interatomic spacings, and the exciton is well described by the theory of the hydrogen atom. The bound-state energies are K=-^-—-
2
(6-6)
with Bohr radius aB given by aR = 4n£oshh2/iie2
(6.7)
where /x is the reduced mass: JU" 1
=m~1 H-m^1
(6.8)
In these expressions eb is the part of the dielectric constant arising from excitations other than the exciton itself. Typically eb is of order 10. In addition, the effective mass in semiconductors of interest is of order one-tenth of the free electron mass. Consequently the exciton Bohr radius, (6.7) is 5 nm or more, while the effective Rydberg constant (coefficient of — \jn2 in (6.6)) is at most some tens of millielectron volts. In particular, the ground-state binding energy |£ x | is much smaller than the semiconductor gap energy £ g . An exciton can move through the crystal as a bound complex. Addition of the kinetic energy of the centre-of-mass motion at wavevector q shows 188
Bulk polaritons that the whole complex has energy g
\ \
(6.9)
where M = me-hmh
(6.10)
In a polar semiconductor, the exciton is a dipole-active mode that couples strongly to light. One consequence is that the onset of optical absorption is at Eg — \EX\ rather than at the direct gap Er The absorbed photon creates an exciton and not a free electron-hole pair, and therefore the energy necessary is reduced by the exciton ground-state energy. In suitable cases such as Cu 2 O a sequence of absorption lines at energies Eg — |£B| for different n is seen on the shoulder of the main absorption threshold at Eg; an example is given in Kittel (1986). A dipole-active mode gives a contribution to the dielectric function. For the exciton, one can take over the classical or quantum-mechanical theory that gives (5.60) for the case of the TO phonon. The resulting expression contains a sum over the contributions from the excited states En in (6.9). The most important is the ground-state term Ex(q), and for the present introductory account we retain only this term. Ex(q) appears in place of the TO photon energy hcoT in (5.60), so the dielectric function is given by e(q, co) = £oo + S/(co2 + Dq2 -co2-
icoT)
(6.11)
whereftcoe= Eg — \EX\ and D = 1/2M. S is the dipole strength of the exciton resonance, and s^ the background dielectric constant. A phenomenological damping parameter F has been introduced. The most important feature of (6.11) is that spatial dispersion enters via the term Dq2; the physical origin of the spatial dispersion is the centre-of-mass contribution to the exciton energy in (6.9). The bulk exciton-polariton dispersion relation is found by solving Maxwell's equations with the use of the dielectric function of (6.11). The result is an obvious generalisation of results obtained in §6.1.1, namely *(q, co)lq2 - e(q9 co)w2/c2^ = 0
(6.12)
This gives a longitudinal mode if e(q, co) = 0 and a transverse mode if the second factor vanishes. For the moment we concentrate on the latter. As explained in §6.1.1, the dispersion curve is most meaningful when damping is neglected. The vanishing of the second factor in (6.12) then gives a quadratic equation in q2. Detailed investigation, left for Problem 6.2, shows that there is one real solution for q for oo< coL, but two for co > coL, where coL(>coe) is the generalisation of the corresponding phononpolariton parameter. The dispersion curve, as sketched in Fig. 6.4, 189
Surface polaritons
resembles that for a phonon-polariton, Fig. 6.1, except that the lower branch bends up for large q. As shown in the insert to Fig. 6.4, the curve may be viewed as resulting from the 'crossing' of the photon dispersion curve with the exciton dispersion of (6.9). An important practical difference is that exciton-polaritons occur near the gap frequency, typically in the visible or near infrared region, whereas phonon-polaritons occur in the far infrared. Fig. 6.4 brings out a major point of interest. Suppose s-polarised light of frequency a> > coL is incident from vacuum on the surface of an excitonic medium. As indicated, two transverse modes, of wavevectors q1 and q2, can propagate in the medium at this frequency. It should therefore be possible to evaluate two transmission coefficients Tx and T2 and one reflection coefficient R to give the amplitudes transmitted into the two propagating modes and reflected into the vacuum. However, Maxwell's equations yield only two boundary conditions. Thus, just as for the dipole-exchange modes discussed in §4.4, an additional boundary condition (ABC) is required. For the dipole-exchange mode the ABC can be found from the microscopic torque equations of motion, since the whole continuum theory may be derived from the microscopic theory. Similarly in the present case the ABC can in principle be derived from the quantum-mechanical equation of motion of the exciton in the presence of the surface. However, since the exciton is an electron-hole pair, what Fig. 6.4 Sketch of exciton-polariton dispersion curves (full lines). Inset shows origin of curve in crossing of photon and exciton curves. The two transverse modes propagating at frequency co > coL are indicated. The longitudinal mode is also shown (broken line).
190
Bulk polaritons is required is in effect the solution of a quantum-mechanical three-body problem. The exact theory appears to be quite intractable, and in practice various ansdtze are used. One popular form, analogous to (4.47) and (4.48), is dP/dz + £P = 0
(6.13)
where P is the excitonic polarisation. Expressions for the reflection and transmission coefficients with this ABC are given by Tilley (1980), for example. An alternative description of the surface is called the dead-layer model. It is assumed that P = 0 in a layer of thickness 2aB at the surface. The boundary conditions that are applied are the usual electromagnetic ones at the two boundaries of the dead layer, together with P = 0 at the interface between the dead layer and the rest of the medium. For p-polarised incident light the situation is somewhat more complicated. In addition to the two transverse modes, a longitudinal mode is excited; as seen from (6.12) its wavevector qL is given by s(qL, a>) = 0. The ABC, (6.13) or whatever else is used, then has two components, so that altogether there are four boundary conditions. These are sufficient to determine the reflection coefficient R, two 'transverse' transmission coefficients 7\ and T2 and a longitudinal' transmission coefficient TL. It might be thought that ordinary reflectivity measurements would discriminate between different ABCs. In practice, however, because of the number of free parameters and variability of specimens, experimental data can be fitted by a range of ABCs. Some discussion of the reflectivity experiments is given by Weisbuch and Ulbrich (1982). The use of Brillouin scattering to investigate exciton-polaritons was suggested by Brenig, Zeyher and Birman (1972), BZB for short. As shown in Fig. 6.4, light of frequency a> > a>L incident on the medium generates two transmitted waves, at wavevectors qx and q2 say. As a simplification, let us first ignore complications due to the absorption of light in the medium and reflection of phonons at the surface. The Brillouin scattering may then be described by the kinematics of simple bulk scattering, as described in §1.4.2. For normal-incidence backscattering, the Brillouin process must scatter a polariton from one of the + q branches onto one of the — q branches. Conservation of frequency and wavevector, (1.52) and (1.53), shows that the frequency shift for Stokes scattering is found from a line of slope vs, where vs is the relevant acoustic phonon velocity. Likewise, anti-Stokes scattering requires a line of slope —v s. These are illustrated in Fig. 6.5. It is seen that for incident frequency co > coL, both Stokes and anti-Stokes lines split into four, corresponding to the four scattering possibilities 1 -• 1', 1 -> 2', 2 -• 1' and 2 -> 2'. 191
Surface polaritons A description of the scattering that includes light absorption and phonon reflection can be given by an extension of the calculation for normalincidence Brillouin scattering outlined in §2.5.2. The details are found in Tilley (1980). It emerges that if the frequency dispersion of the dielectric function is not too great, each scattering peak has the Dervisch-Loudon lineshape. The theory also gives expressions for relative intensities of the different lines. Resonant Brillouin scattering has been observed in a wide range of polar semiconductors. An example is shown in Fig. 6.6. It should be noted that the differential cross section changes very rapidly with incident photon energy, and for this reason these experiments are always performed with a tunable dye laser as source. It is seen from Fig. 6.6 that, as expected from the kinematics of Fig. 6.5 only one Stokes peak is seen for a> < a>L, whereas a multiplet is observed for ftco = ha>L + 0.25 meV. For larger co, the multiplet is washed out by the increasing opacity broadening. The peak positions in results like those shown in Fig. 6.6 can be fitted Fig. 6.5 An illustration of the kinematics of Brillouin scattering via exciton-polaritons [after Weisbuch and Ulbrich 1982].
*r
192
Bulk polaritons
by the exciton-polariton model of Fig. 6.5, and by this means the polariton parameters are deduced. An example of such a fit for GaAs is shown in Fig. 6.7. Theoretical fits like that shown in Fig. 6.7 are given by the simple kinematics of Fig. 6.5; they do not test the full theory for lineshapes and relative intensities. The full theory has been tested to a limited extent by So et al. (1981). Since the incident and scattered photon energies are close to the band gap (that is the significance of the word resonant) the refractive index changes very rapidly with photon energy. Consequently the simple expression for the Stokes :anti-Stokes ratio given in (1.54) no longer applies. The results of Fig. 6.6 serve as an illustration. For a>L the anti-Stokes intensity is greater than the Stokes, and the ratio of the intensities clearly varies with photon energy. Thus the frequency dependence of the Stokes:anti-Stokes ratio may be used for comparison of theory and experiment, as shown in Fig. 6.8. The two theoretical curves shown correspond to slightly different ways of extending the theory of §2.5.2. In the space available here, it has not been possible to give more than a brief introduction to exciton-polaritons. A much fuller account is given by Weisbuch and Ulbrich (1982). Fig. 6.6 Resonant Brillouin scattering in GaAs at 12 K for incident light frequencies around coL (hcoL = 1.60 eV). Stokes lines are to the left [after Weisbuch and Ulbrich 1982^
AJ^^A^tr^^Wr
1|B^^^^J
Sk
+0.25
-0.30
-0.55
-^W -0.2
I 0
V^_ 1.40 +0.2
A£[meV]
193
Surface polaritons 6.2 Surface polaritons: single-interface modes 6.2.1 Basic properties
It has been known for many years that an electromagnetic wave can propagate along an interface between two media of which at least one is dispersive. In fact the calculation to be presented in this section for a mode at the interface between two isotropic media appears as a problem in Landau and Lifshitz (1971). Interest increased greatly from the late 1960s, when ATR was established as a viable technique for experimental study of surface polaritons. Detailed reviews of that work are given by Otto (1974 and 1976) as well as in several articles in Burstein and De Martini (1974). For a recent account of the whole subject, the reader may refer to Agranovich and Mills (1982). In this chapter, as elsewhere in the book, we shall give theory first, followed by a discussion of experimental results. Thus this section and the next are theoretical; the ATR experiments are the subject of §6.5. We consider the plane interface between two semi-infinite dielectric media; as shown in Fig. 6.9 we take the z axis normal to the interface and the x axis as the direction of propagation of the mode. Explicit Fig. 6.7 Measured Brillouin shifts as a function of incident photon energy in GaAs. The full lines are theoretical curves deduced from the model indicated in Fig. 6.5 [after Weisbuch and Ulbrich 1982].
+ 0.5
-0.2
0 AE (meV)
194
+0.2
Surface polaritons: single-interface modes calculation shows that there is no surface mode with the E field in the y direction, so we therefore take E in the xz plane: E = (Elx, 0, Elz) exp(iqxx - icot) exp(i
(6.14)
in medium 1, with a similar form in medium 2. We take a single frequency a>, and since boundary conditions will be applied on the whole plane z = 0 the wavevector qx must be the same in both media. Clearly qx is the propagation vector, and one of our aims is to find the dispersion equation
For the mode to be localised, E must decrease with distance from the interface. Thus lm(qlz) > 0 and lm(q2z) < 0. In order for (6.14) to satisfy Fig. 6.8 Stokes:anti-Stokes intensity ratio versus wavelength for resonant Brillouin scattering via exciton-polaritons in Cu 2 O. Full line, theory of So et al. (1981a); broken line, theory of Tilley (1980) [after So et al. 1981b]. 1.4-1
1.1-
\ 0.8-
0.5576.6
577.0 577.4 Wavelength (nm)
577.1
Fig. 6.9 Notation for surface-polariton calculation. Medium 1 occupies the half-space z > 0, and medium 2, z < 0.
z= 0
195
Surface polaritons Maxwell's equations in both media, the wavevectors must satisfy o)2/c2
i=l,2
(6.15)
In particular, this means that when sx and e2 are real, the localisation requirement is ql > e^/c2. The equation V • D = Ogives the ratios of the field amplitudes in (6.14): qxEiX + qizEiz = 0
1=1,2
(6.16)
The amplitudes in the two media are related by continuity of the tangential component of E and the normal component of D: Elx = E2x
(6.17)
£iElz = s2E2z
(6.18)
Equations (6.16) to (6.18) are four homogeneous equations in the four field amplitudes, and the solvability condition is the required dispersion relation. Equations (6.16) and (6.17) give Eiz in terms of Elx, then substitution in (6.18) gives the dispersion relation in the implicit form qiz
q2z
(6.19)
£2
Substitution from (6.15) leads to the explicit result
alA^
C2 £i + £ 2
(6-20)
As for bulk polaritons, we start by neglecting damping, so that e1 and e2 are both real. It then follows from (6.15) and the localisation requirement that qlz and q2z are both pure imaginary, and we write *iz = i*i (6-21) q2z=-iK2 (6.22) where the signs incorporate the localisation condition. Equation (6.19) now shows that for the surface polariton to exist, e1 and s2 must have opposite signs: £l£2<0
(6.23)
Since the right-hand side of (6.20) must be positive, we have the further condition e1+e2<0 (6.24) Equations (6.23) and (6.24) determine the frequency intervals in which the surface polaritons occur. In many experimental situations, one of the dielectric functions is frequency-independent and positive, as is the case if medium 1, say, is vacuum. In that case, (6.23) and (6.24) show that the other dielectric constant, s2 say, has to be negative, e2 < — £ t . Medium 2 is then called the surface-active medium. 196
Surface polaritons: single-interface modes As a concrete example, we consider the interface between vacuum, with ex = 1, and a polar medium described by the dielectric function of (5.60) (see also Fig. 5.9). Equation (6.24) then shows that the surface polariton occupies the frequency interval coT<(Ds
(6.25)
where cos is given by fi2(G>s)=-l
(6.26)
or explicitly ^s2 = (eoo^ + a)|)/( £oo + l)
(6.27)
As anticipated, the frequency interval is within the stop band for bulk polaritons. The dispersion curve is illustrated in Fig. 6.10; the asymptotic values are qx -• co/c
\qiz\ -• 0
a s co -* coT
(6.28)
qx -• oo
\q iz\ -• oo
a s co -+ cos
(6.29)
and These limiting values for \qiz\ show that the surface polariton is weakly localised at the low-frequency end, and strongly localised at the high-frequency end. For a second example, we take e1 = l and s2(co) in the plasma form of (5.51), again with damping neglected. The equations replacing (6.25) and (6.27) are 0 < co < cos
(6.30)
G>| = a # 2
(6.31)
and The dispersion curve is illustrated in Fig. 6.11; equations (6.28) and (6.29) apply on the understanding that coT is replaced by 0. The surface modes at the interface between a vacuum and a medium with dielectric properties specified by (5.52) are discussed by Palik et a\. (1976). Equations (6.16) to (6.19) may be used to find the relative amplitudes and phases of the electric-field components in the two media, and the magnetic-field components can be derived from Maxwell's equations in the usual way. Furthermore, Poynting's theorem can be applied to find the directions and magnitudes of the energy flow in the media; details are given by Nkoma et al. (1974). The results are illustrated in Fig. 6.12 for the case when 81>0. 197
Surface polaritons Fig. 6.10 (a) Surface polariton dispersion curve for a Ga As/vacuum interface. Numerical values for GaAs are quoted in Fig. 5.9; with these values cos/coT= 1.081. {b) Dispersion curve of (a) on an expanded scale. 2.00
1.08T
1.00
198
1.50
2.00
2.50
3.00
Surface polaritons: single-interface modes Fig. 6.11 (a) Surface plasmon-polariton dispersion curve for a metal/vacuum interface. The graph applies for any metal with the dielectric function of (5.51). (b) Dispersion curve of (a) on an expanded scale. 3.0
2.0
1.0
2.0
1.0
3.0
0.6-
0.4-
0.2-
1.0
2.0
3.0
199
Surface polaritons 6.2.2 Electrostatic approximation The frequency a>s for which qx -> oo in dispersion curves such as those of Figs. 6.10 and 6.11 is of particular interest. For example, in electron energy-loss experiments on metals the momentum transfer is relatively large, and a loss peak is observed at a>s. The condition qx » a>/c means that the wavelength 2n/qx of the wave along the surface is much smaller than the free-space wavelength Inc/co. Retardation effects, that is, the nonzero propagation time of an electromagnetic signal, are therefore unimportant, and the asymptotic wave with co ~ cos and qx » a>s/c is called an electrostatic surface wave. It is seen from (6.20) that the (fixed) frequency of the electrostatic surface wave is given by 81+£2=0
(6.32)
with explicit expressions for the vacuum-bounded polar dielectric and plasma given by (6.27) and (6.31) respectively. In view of the importance of the result, we now derive (6.32) by means of an explicit electrostatic approximation. The calculation is the analogue of that presented for the magnetostatic surface wave in §4.2. In the electrostatic approximation, media 1 and 2 are described by potentials Kx and V2 satisfying V 2V t = 0
i = 1,2
(6.33)
Fig. 6.12 Electric and magnetic field amplitudes and Poynting vector for an interface with Ei > 0 (upper medium) and s2 < 0 (lower medium), {a) Electric field pattern in the xz plane. The field magnitudes are oscillatory in the x direction but decrease exponentially away from the boundary, {b) An analogous diagram of the magnetic field pattern in the xy plane, (c) Poynting vector pattern close to the boundary. The energy flow in medium 1 is greater than that in medium 2, giving a net flow of energy to the right when qx is positive [after Nkoma et al. 1974].
— \
\
f
(a)
yi
t
t
t
t
!
t
(b)
I
I — \
\ 200
—
\
Surface polaritons: single-interface modes A plane wave localised at the interface and travelling in the x direction is described by Vx = V10 Qxp(-Kxz) exp[i(qxx - cot)~]
(6.34)
with a similar expression for V2. Substitution into (6.33), which involves only spatial derivatives, gives Kf
= ql
i=l,2
(6.35)
Thus the potentials are Vx = V10 cxp(-qxz)
exp[i(qxx - cot)]
V2 = V20 exp(qxz) exp[i(gxx - cot)]
(6.36) (6.37)
with fields E = — W given by E1 = ( - i « x , 0 , ^ ) F 1
(6.38)
E2 = ( - i ^ , 0 , - ^ ) K 2
(6.39)
Equations (6.36) and (6.37) have the property, which they share with magnetostatic waves in the Voigt configuration, that the decay constant normal to the interface is equal to the wave vector along the interface. Obviously this follows from the fact that the potential satisfies Laplace's equation (6.33). It is seen further from (6.38) and (6.39) that the field components Ex and Ez in either medium are equal in magnitude but 90° out of phase. This is a special case of the general surface-polariton field patterns illustrated in Fig. 6.12. In order to derive the frequency of the electrostatic surface wave, we apply standard boundary conditions to (6.38) and (6.39). Continuity of Ex gives V10 = V20
(6.40)
and then continuity of Dz gives ^(1x^10=
-^iClxV20
(6-41)
In view of (6.40), the above equation reduces, as expected, to the result previously derived in (6.32). The calculations we have presented in this section apply to linear media, in which the dielectric functions are independent of field strength. Laser sources of sufficient intensity are nowadays available for nonlinearities in the dielectric response to be important. The inclusion of nonlinearities leads to a number of new effects, for example the appearance of a 'self-guided' TE mode for incident power above a threshold intensity. Self guiding is more easily understood by reference to two-interface modes, which are the subject of the next section, so the discussion is deferred until §6.4. 201
Surface polaritons 6.2.3 Anisotropic media An important extension of the results so far derived is to the case when one of the media is anisotropic. A brief history with references is given by Mirlin (1982); convenient reviews are given by Borstel et al. (1974) and by Borstel and Falge (1977 and 1978). Here we shall simply quote a result that will be useful in later sections. We continue to follow the notation of Fig. 6.9, so that (x, y, z) are 'surface-related' axes. Now, however, we suppose that medium 1 is isotropic but medium 2 is anisotropic, so that its optical properties are characterised by a dielectric tensor s2(co). This tensor has principal axes, (x', / , z') say, and considerable algebraic complexities arise because in general these axes are at an arbitrary orientation with respect to the surface-related axes (x, y, z). As before, we consider a mode propagating in the x direction, q,, = (qx, 0). It is sufficient for a simple illustration to consider the special case when the y axis is a principal axis of s2(co), y — y' say. In this case the boundary conditions at the surface do not mix s- and p-polarisations, and we choose to deal with the latter, E = (Ex, 0, Ez). With these simplifications, it can be shown that the dispersion equation for the surface polariton is
rf-^%^ C
CyO.
(6-42)
£1
Here e'x and ez are the principal values of s2 in the x'z' plane, while ezz is the appropriate component of s2 in (x, y, z) axes. The reader is invited in Problem 6.4, to verify the special case of (6.42) when (x, y, z) and (x\ / , zf) axes coincide. One application of these results is to ferroelectrics. In §3.6 we dealt with surface magnons in the transverse Ising model, which describes hydrogenbonded ferroelectrics like KDP. For long wavelength modes, coupling to the electromagnetic field becomes important. Many ferroelectrics, both hydrogen-bonded and the other main class, displacive, can be characterised in the simplest approximation by a uniaxial dielectric constant of the form /e±
0
0'
£ = J 0 s± 0 I
(6.43)
\ 0 0 in terms of principal axes. The component £(| has a significant frequency and temperature dependence arising from the phase transition, while s± can be treated as constant. In a displacive ferroelectric, such as BaTiO 3 , the transition at a temperature Tc from paraelectric to ferroelectric phases consists of a relative displacement of the positive and negative ions in the 202
Surface polaritons: single-interface modes lattice. This displacement has the same symmetry as a long-wavelength optic phonon. Thus the restoring force for the optic phonon decreases as T -> Tc, and the phonon frequency coT decreases. This means that e^ can be represented by the TO phonon form of (5.60), with a temperaturedependent resonant frequency. Near to Tc, however, the damping becomes important and other mechanisms come into play with the effect that £|| has a more complicated form. It is clear that in principle the surface-polariton properties may be used to investigate the behaviour of g||. Some discussion of the possibilities is given by Cottam et al. (1984), but there have not yet been any experimental investigations.
6.2.4 Charge-sheet modes In §5.3.4 it was mentioned that in suitable circumstances very thin sheets of mobile electrons can be trapped at an interface, and that these sheets can be described as a 2D electron gas. As we now show, a surface-polariton-like mode is predicted to propagate parallel to the charge sheet, with the amplitudes localised at the position of the sheet. The derivation of the dispersion relation is a straightforward extension of that given in §6.2.1 for the ordinary surface polariton. We use the axes of Fig. 6.9 and the notation of §6.2.1, but we now suppose that a 2D charge sheet of particle density v m~ 2 is localised at the interface z = 0. Equations (6.14) to (6.17) still apply, but (6.18) must be replaced. It is equivalent to the continuity of the field component Hy; since there is now a sheet of mobile charge at the interface the correct form is H2y-Hly=jx
(6.44)
where j x is the current density in the charge sheet. As in the 3D plasma, j x is driven by the alternating electric field at the interface, Elx, which from (6.17) is equal to E2x. We ignore collisions in the electron gas, so we can write from the 2D equivalents of (5.43) and (5.47) j x = ive2Elx/mcc>
(6.45)
Substitution of X, from (6.45) in (6.44) gives the relation that replaces (6.19), namely ^1 + ^ = ^ Kx
K2
CO2
(6.46)
where KX and K2 are still given by (6.21) and (6.22). The frequency Q s is given by Q s = Ve2/eomc
(6.47) 203
Surface polaritons In the electrostatic limit, it follows from (6.15), (6.21) and (6.22) that KX and K2 are both replaced by qx. Equation (6.46) then simplifies to o) = inscqx/(si+s2)y/2 (6.48) The dispersion relation (6.46) was first derived by Nakayama (1974) by the method described here. Equation (6.48) was derived earlier by Stern (1967), who used linear-response theory rather than simple dynamics to describe the electron sheet. An important property of (6.46) and (6.48) is that a localised mode can propagate even when £x and e2 are both positive. This is in striking contrast to the ordinary polariton, for which e1 and e2 must have opposite signs, (6.23). The dispersion curve for ex = e2 is shown in Fig. 6.13. There has not been a great deal of experimental work on these plasmon modes of a single charge sheet. One exception is the study by Batke and Heitmann (1984) of space-charge layers produced at a semiconductorinsulator interface by application of a voltage to a superposed electrode. They couple far-infrared radiation to the plasmon by means of a grating, as will be discussed briefly in §6.5.1. Their transmission spectra show clear evidence of the charge-sheet plasmon. 6.2.5 Surface exciton-polaritons In §6.1.2 we saw how the centre-of-mass momentum of the exciton leads to significant spatial dispersion in the dielectric function. We now consider how the calculation of surface-polariton properties, as given in §6.2.1, is modified by spatial dispersion.
Fig. 6.13 Charge sheet dispersion curve corresponding to (6.46) for ex = e2, both taken positive. 10 r
Cf
100
204
200
Surface polaritons: single-interface modes In the absence of spatial dispersion, the surface polariton is found in some frequency range within the bulk-polariton stop band; examples are given in (6.25) and (6.30). As shown in Fig. 6.4, however, the bulk exciton-polariton has no stop band; the branch labelled 2 there has the large-g behaviour co ~ Dq2 because of the spatial dispersion. The situation is closely analogous to that described for dipole-exchange modes in §4.4. In the absence of exchange effects, pure magnetostatic surface modes propagate, as seen in §4.2. The inclusion of exchange, which is necessary for larger q values, means that the magnetostatic mode becomes 'leaky', and radiates into the bulk mode with which it is degenerate in frequency. Similarly in the present case there is no uncoupled surface excitonpolariton; instead we have a leaky surface mode. A theoretical treatment of the surface exciton-polariton analogous to that given in §6.2.1 for the phonon- or plasmon-polariton was given for the first time by Maradudin and Mills (1973). In their calculation the medium is described by the isotropic dielectric function of (6.11); we summarise it with reference to Fig. 6.9. It is assumed that the mode is p-polarised, so that in the non-surface-active medium the electric field is given by (6.14) with wavevector qlz satisfying (6.15). In the excitonic medium, however, q2z is found as a function of qx and co from (6.12). In the absence of damping, (6.12) has one real root for q as a function of co, namely branch 2, for co < coL, and three real roots for co > coL. There is no frequency range in which only imaginary roots are found. Likewise, when (6.12) is regarded as giving q2z as a function of qx and co, the root corresponding to branch 2 is always real. With damping present, this root is off the real axis, but for a realistic value of the damping it is still predominantly real. The other two roots for q2z are predominantly imaginary for small co, and real for larger co. As in the calculation of reflection and transmission of p-polarised light, mentioned in §6.1.2, the electric field in the excitonic medium has three components, so that the analogue of (6.14) is E= t
Plx, 0, E\z) exp(i
(6.49)
Here a labels the three solutions of (6.12) for q2z. The amplitude E2z is related to Ea2x by V • D = 0 for the transverse modes, and by V x E = 0 for the longitudinal mode. In all, then, four field amplitudes have been defined, namely Elx and Ea2x for a = 1, 2, 3. Four linear homogeneous equations in these amplitudes are found from the two electromagnetic boundary conditions and two ABCs resulting from (6.13) or whatever is used instead. The condition for these equations to have a solution gives the dispersion 205
Surface polaritons
relation qx((o). As anticipated, it is found that even in the absence of damping qx is complex, so that the surface wave is attenuated. The attenuation is due to energy leakage into the bulk branch 2. With the damping parameter in (6.11) included, the attenuation of the surface mode increases. As was seen in §6.1.1, the definition of a dispersion curve becomes ambiguous in the presence of attenuation, since the dispersion relation is then between two complex variables a> and qx. Nevertheless, for small attenuation a dispersion curve can be a useful guide to the eye, and an example of a calculated curve is shown in Fig. 6.14. The calculation outlined here assumed an isotropic medium; many excitonic media are anisotropic, and the calculation must be appropriately generalised. The experimental study of surface exciton-polaritons began with studies of the ZnO surface by Lagois and Fischer (1976) and De Martini et al. (1977); a detailed account of this and later work on ZnO is given by Lagois and Fischer (1982). The former group used the method of attenuated total reflection (ATR), and a discussion of their results is deferred to §6.5 which gives a detailed account of ATR. The latter group generated the surface exciton-polariton by means of frequency-doubling the light from an intense dye laser. The principle of the method is sketched in Fig. 6.15(a). ZnO Fig. 6.14 Calculated dispersion curve of surface exciton-polaritons for complex frequency GO and real wavenumber qx. The calculation was done for a ZnO exciton using the ABC P = 0, which is a special case of (6.13) [after Lagois and Fischer 1978],
206
Surface polaritons: single-interface modes does not have a centre of inversion symmetry, so the incoming light of frequency v and in-plane wavevector kx can generate a surface mode of frequency co = 2v and wavevector qx = 2kx. Detection of the surface mode was simple in principle. As will be seen at the end of §6.5 in a little more detail, roughness on the specimen surface destroys the in-plane translational invariance, and allows the surface mode to radiate. Wavevector qx can Fig. 6.15 (a) Sketch to show principle of nonlinear optical excitation of surface excitonpolariton on ZnO used by De Martini et al. (1977). (b) Experimental results for real and imaginary part of qx = q'x + \q"x obtained by De Martini et al. Curves are theoretical with parameters 6^ = 6.15, e(0) = 6.172, ha>T = 3.421 eV and fcr = 0.25 meV [after De Martini et al. 1977].
(a)
12.5
3.422
2.0
2.2 5
2.4
1
Wavevector ^ ( 1 0 cm" )
207
Surface polaritons
be varied by varying the angle of incidence 6. For given excitation frequency v and mode frequency a> = 2v the wavevector qx corresponding to peak intensity gives a point on the dispersion curve, while the linewidth in qx is a measure of I m ^ ) . The results obtained are shown in Fig. 6.15(fo). It may be noted that De Martini et al. analysed their results using a model that lacks spatial dispersion; the parameters quoted in the caption to Fig. 6.15 correspond to the theoretical expression of (5.60). Later, however, Fukui et al. (1980) showed that a good account can also be given when spatial dispersion is included. Most recent experimental and theoretical work on surface excitonpolaritons has been concerned with either ATR or nonlinear generation, although Brodin et al. (1984) carried out luminescence studies on ZnTe. Fukui and Tada (1982) proposed that as an extension of the method described in §6.1.2 surface exciton-polaritons could be probed via resonant Brillouin scattering. The principle is sketched in Fig. 6.16. Fukui and Tada give a detailed theoretical analysis, which is a generalisation of that outlined in §6.1.2, but no experimental results have yet been reported. 6.3 Surface polaritons: two-interface modes
We now extend the calculations of the previous section to a geometry with two plane interfaces. The notation is indicated in Fig. 6.17. Fig. 6.16 Principle of resonant Brillouin scattering experiment proposed by Fukui and Tada (1982). Possible scattering processes are indicated [after Fukui and Tada 1982].
208
Surface polaritons: two-interface modes
As before, we take the z axis normal to the interfaces, which are situated at z = 0 and z= —L. The x axis is taken along the direction of propagation. As was discussed in §2.3 for surface elastic waves, for sufficiently large L independent surface polaritons propagate on each interface. If we imagine L decreasing, thefieldsof the two surface polaritons propagating for a given qx start to overlap, and the degeneracy in frequency between them is lifted. Thus we expect to see two modes, essentially a bonding and an antibonding combination. The behaviour for magnon-polaritons is rather different (see §6.7), but even in the case of phonon- or plasmonpolaritons there are other possibilities from that just described. In the surface-polariton calculation of §6.2.1 q2z was constrained to be imaginary in order that the polariton field should tend to zero as z -• — oo. In the present case, qlz and q3z must be imaginary (in the absence of damping), but q2z can be either imaginary or real. Imaginary q2z corresponds to the description just given, where the surface polariton on one interface is perturbed by the presence of the other interface. For real q2z, on the other hand, the z-dependence of the fields in medium 2 is oscillatory, so that medium 2 is behaving like a waveguide. For this reason the mode is called a guided-wave polariton.
The previous paragraph was concerned, implicitly, with p-polarised modes. Once the possibility of guided-wave polaritons is recognised it is necessary to discuss also s-polarised guided-wave modes. It turns out that these can occur, and indeed so far they have been the modes of most importance in applications of dielectric waveguides. We first derive the governing equations for the modes, both surface-polariton and guided-wave type. The calculation is a straightforward extension of that given in the previous section, and some of the
Fig. 6.17 Notation for two-interface calculations.
z= 0
-L
209
Surface polaritons equations presented there still apply. We write the E fields in the three media as E = (Elx, 0, -qxEJqlz)
exp(iqxx - icot) exp(i^flzz)
z> 0
(6.50)
E = {a exp[i2z(z + L/2)] + b exp[-i 2z (z + L/2)], 0, - qxa exp[i2z(z + L/2)]/q2x + qxb exp[-ig 2 z (z + L/2)]/q2z} x exp(iqxx — icot)
0> z> —L
(6.51)
E = (£ 3 x , 0, - qxE3x/q3z) exp(iqxx - icot) exp[i3z(z -f L)]
z< - L (6.52)
Here (6.16) has been applied to relate the z and x components in each medium. The localisation condition is Im(g lz ) > 0 and Im(g 3z ) < 0, which requires ql>efo2/c2
i=l,3
(6.53)
The four amplitudes Elx9 a, b and E3x are related by four boundary conditions, two at each interface. They take the form af+bf-'=Elx fi2(^-V1)/«2, qf'1-bf=E3x
(6.54) = Ci£i7«i.
(6-55) (6.56)
*2W~l ~ bf)/q2s = e3E3x/q3z
(6.57)
/=exp(i
(6.58)
where Equations (6.54) to (6.57) are four homogeneous equations in the amplitudes, and the solvability condition gives the dispersion equation. It is readily seen that the general form is
± ^
(6.59)
This is somewhat too general for comfort, and we therefore specialise to the symmetric geometry, sx =s3. The localisation condition becomes qlz= — q3z = \K1
wit\\K1>0
(6.60)
and (6.59) is seen to have two solutions exp(i4 2z L)= ±(elq2z + ie2K1)/(e1q2z-i£2Kl)
(6.61)
Equation (6.61) applies equally to surface-polariton and guided-wave modes. For surface-polariton modes, q2z is pure imaginary, g 2z = k 2 . The two solutions of (6.61) can then be reorganised into the forms
210
e2/e1 = —(K2/K1)
tanh(^K:2L)
e 2 / fil = -(KjKi)
coth(^c 2 L)
(6.62) (6.63)
Surface polaritons: two-interface modes As L -• oo, the hyperbolic functions in (6.62) and (6.63) tend to one, so that both equations are asymptotically the same as the dispersion relation for the single-interface mode in the form given by (6.19), (6.21) and (6.22). Thus, as expected, (6.62) and (6.63) do describe the bonding and antibonding combinations of the surface-polariton modes. The general form of the solutions is illustrated for two different thicknesses of LiF slab in Fig. 6.18. So far in this section it has been assumed that all three dielectric constants are real, so that as usual in the discussion of dispersion relations damping has been neglected. Fukui et al. (1979) and Sarid (1981) carried out numerical studies of the effect of including damping. In both cases, the starting point was the dispersion equation, (6.59) in general, or (6.62) and (6.63) for a symmetric case. Fukui et al. took qx real and calculated the imaginary part of co, while Sarid took co real and calculated lm(qx). These different assumptions correspond, ultimately, to different experimental arrangements. For example, real co corresponds to a wave launched along the film at that frequency, and 1/Im(gx) is then a measure of the decay length of the wave along the film. Both Fukui et al. and Sarid considered the particular case of metal films and plasmon-polaritons, so that the dielectric function of the film had the form of (5.51). The important result obtained by both groups concerns the higher-frequency mode. Sarid's
Fig. 6.18 Two-interface surface polaritons for a LiF slab bounded by vacuum. 1, qTL = 0.1 (L = 0.5 /an); 2, qTL = 0.2 (L=1.0/mi); 3, qTL>2 (L> 10 /mi). Here qT = coT/c. The full curves represent the calculation with retardation, the broken curves without retardation [after Bryksin et al. 1974].
1.5
0
1
2
10
211
Surface polaritons result is that its decay length increases as the film thickness decreases, and correspondingly Fukui et al. find that the lifetime 1/Im(a>) increases as the thickness decreases. The mode is therefore called the long-range surface plasmon (LRSP). In the LRSP for a film between identical dielectrics the electric-field pattern is antisymmetric across the film, while in the lower-frequency mode the electric-field pattern is symmetric. The long decay length of the LRSP may be seen as arising from the fact that as the film thickness decreases a smaller proportion of the mode energy is transported within the film. Experimental confirmation of these results has been obtained by a number of groups, one of the earliest to appear being that by Craig et al. (1983). A related experimental curve obtained by ATR is shown subsequently. In applications to nonlinear optics it is helpful to use long-range modes, so the LRSP is likely to be of considerable technical importance. As stated, the earlier work was concerned with surface plasmonpolaritons. The extension to surface phonon-polaritons was carried out by a number of authors; a good account is given by Wendler and Haupt (1986a), who have also discussed (Wendler and Haupt 1986b) the surface plasmon-phonon-polariton found in a doped semiconductor when the plasma frequency is close to the TO phonon frequency. We now turn to guided modes in p-polarisation. For real q2z, the two forms of (6.61) are fiz/fii = (2z/*i) tan(ig 2z L) HlH = -(qijKy) cot£q2zL)
(6.64) (6.65)
as can also be seen directly from (6.62) and (6.63). The numerical, or graphical, task of solving these is similar to that of finding the odd and even parity modes in a quantum-mechanical square well, or of finding the dispersion curves for the Love waves mentioned in §2.3.1. For small L, there is only a small number of modes; the number increases as L increases until for large enough L the guided-wave modes form a quasi-continuum, which in the limit becomes the bulk continuum. The dispersion curves are illustrated for a moderate value of L in Fig. 6.19(a). Comparison with Fig. 6.10(a) shows how the curves occupy the parts of the a>qx plane that ultimately as L -• oo are occupied by the bulk continuum. The calculation for the s-polarised guided-wave modes is similar to that for the p-polarised modes, and the details are left for a problem. It emerges that the equations for p-polarisation can be converted to those for s-polarisation by the substitutions £2
(6.66)
Surface polaritons: two-interface modes (6-67) Thus the general result, taken over from (6.59), is (6.68) while for a symmetric geometry the equations are (6.69) (6.70) The corresponding dispersion curves are shown in Fig. 6.19(b). We should make an elementary comment which is implicit in the formalism presented. In applications, dielectric waveguides are usually fabricated from materials which all have positive dielectric functions. In the geometry discussed, Fig. 6.17, an essential condition for guided waves to propagate is that q2z should be real while qlz and q3z are imaginary.
Fig. 6.19 Dispersion curves of guided-wave polaritons in a film placed in vacuum, {a) ppolarised, (b) s-polarised [after Ushioda and Loudon 1982]. (a) TM modes
(D = Cqx/€2((i)Y'
(b) TE modes
213
Surface polaritons
In view of (6.15), this requires simply that s2 is the largest of the three dielectric constants, or in terms of the refractive indices ni = ej /2 , n 2 >max(n 1 ,n 3 ) (6.71) This is the waveguiding condition for the geometry of Fig. 6.17. The guided waves that have been described are of importance in engineering applications; further details are given in the relevant chapter of Marcuse (1982). It may be noted that engineers frequently use the notation qx =fico/c.Comparison with the equation q = rjco/c for a bulk wave in a medium of refractive index Y\ means that p is described as the effective refractive index. Theoretical and experimental results may be presented as a graph of /? versus a>; this amounts to a dispersion graph of the kind used here with the horizontal and vertical axes interchanged. 6.4 Surface polaritons: nonlinear effects
The results presented so far in this chapter have been derived on the assumption that the dielectric responses of all the media involved are linear. This assumption breaks down for sufficiently intense sources and media with strong nonlinear response. The inclusion of nonlinearities leads to a range of new physical effects, and in addition all methods of signal processing rely on the exploitation of nonlinearities. Thus both because of the physics involved and because of potential applications, the properties of nonlinear surface and guided waves are of great interest. At the time of writing the subject is in a very early stage of development, and the full range of possibilities is not mapped out. In this section, therefore, we give some introductory comments and present an exact calculation for the simplest case of self guiding. The interested reader may refer to the much fuller introduction to the subject given by Stegeman et al. (1986) and to the literature review contained in Wendler (1986). When the dielectric functions are linear the governing equations for the electromagnetic field in the presence of boundaries are a set of linear differential equations and boundary conditions. In that case a number of powerful and general results apply. In particular, the principle of superposition holds, so that the sum of two solutions is itself a solution. Furthermore, it is possible to find for a given frequency a complete set of solutions which have the property that an arbitrary solution at that frequency can be expressed as a linear combination of functions of the complete set. For the two-interface geometry of the previous section, a complete set comprises the surface-polariton modes and the guided-wave modes together with the 'radiation modes' in which the wavevector qz is real in media 1 and/or 3. In the presence of nonlinearity, these results no 214
Surface polaritons: nonlinear effects longer apply, so that a general discussion is much more difficult and we do not attempt anything of the sort. Nonlinear optics proceeds by expansion in a Taylor series of the nonlinear polarisation P N L induced by one or more intense electromagnetic fields. The lowest-order effect involves a susceptibility # (2):
^L2 = I * 8 W j,k
(6.72)
This generates wave-like solutions with sum and difference frequencies coa ± ojb, where a>a and cob are the frequencies of the fields E a and E b . The tensor x\]i *s °f third rank, and the number of nonvanishing elements depends in the usual way (Nye 1985) on the crystal symmetry of the medium. The next highest order involves three E components:
P?L3 = I zSi£J£j£f
(6.73)
j,k,l
One reason why effects depending on the fourth-rank tensor %(3) are of interest is that if all three amplitudes E a , E b , E c are at the same frequency co, then P N L 3 contains a component at frequency co, since co — co + co is among the frequencies produced. Thus %(3), in contrast to x{2\ can lead to nonlinear effects at a single frequency co. One such effect is the appearance of an intensity-dependent refractive index; in terms of the dielectric function we write E -• s + a|£| 2 . The example to be considered here is the self guiding of a TE mode at the interface between two media, one of which has such an intensity-dependent dielectric function. The original calculation was by Kaplan (1977). We use the notation and axes of Fig. 6.9, but it is assumed that ex is intensity-dependent: £1-»e 1+a1|E|2
(6.74)
while £2 is independent of intensity. We show that a TE mode can be 'self-guided' and propagate along the interface provided that OLX is positive and the intensity is sufficiently large. The physical reason follows from the comment made at the end of §6.3, that the waveguiding condition for a film is n2 >max(rc 1 , n3). A sufficiently intense wave propagating in medium 1 near the interface produces an increase in the refractive index near the interface because of the second term in (6.74). If this local' value is larger than the refractive index of medium 2, then the wave can be self guiding. Roughly, then, the intensity condition for guiding is Sx + OLi\E\2 > £2.
We assume we are dealing with a single-frequency mode propagating in the x direction, so that all amplitudes vary as exp(iqxx — icot), and as stated we consider a TE mode, so that the E field is in the y direction. 215
Surface polaritons The wave equation in medium 1 is ^
- (ql - cohjc2 - co^El\/c2)Ex = 0
If this is multiplied by dEJdz integral
(6.75)
it can be integrated to yield the energy
\ We are assuming here, as will be justified subsequently, that we can find a solution in which E1 is real. The constant of integration K has the same value for all z. However, we are seeking a solution localised at the interface, in which E1 -• 0 and dEJdz -* 0 as z -» oo. Substituting these asymptotic values, we see that K = 0. The solution of (6.76) is therefore
"kf
where, as in §6.2.1 K\
= q2x- co2ejc2
(6.78)
The integral can be evaluated in terms of hyperbolic functions, and the solution of (6.77) is Ex = (2/a1)1/2(cfc1/co) sechC/c^z - z 0 )]
(6.79)
Here z 0 is a constant of integration, which has yet to be determined. The lower medium is linear, and for a guided wave we have the solution familiar from §6.2.1 E2 = A exp(/c2z) exp(iqxx — icot)
(6.80)
where K2 is given by (6.78) with s2 in place of £x. The boundary conditions are that E and dE/dz are continuous, since the latter is proportional to Hx. Thus (2/a1)1/2(c/c1/co) sech^qzo) = A 1/2
(2/a1) (c/cf/co) tanh(K1z0) sech^Zo) = K2A
(6.81) (6.82)
Elimination of A yields jq tanh(fqz 0 ) = K2
(6.83)
Let us take stock of these equations. We may take the frequency co as given, so the unknowns are qx, A and z 0 . Unhappily we have only the two equations (6.81) and (6.82), so clearly something is missing. The clue is that we are dealing with a nonlinear system, and the form of the solution therefore depends in a nontrivial way on the wave intensity, or equivalently on the field amplitude. This is already clear from (6.79), where the amplitude has a value that depends on the other parameters *q and z 0 216
Surface polaritons: nonlinear effects as well as on the nonlinear coefficient oc1. We must therefore introduce the power explicitly by means of Poynting's theorem. The instantaneous power per unit length of y axis is
-I
(E x H) dz
(6.84)
and the intensity is given by the time-average < ^ > . Evaluation from the field expressions of (6.79) and (6.80) is elementary, if a little laborious. The instantaneous power &> has both a z and an x component. However, the former contains a factor cos(^x — cot) sin(qxx — cot) (recall that it is necessary to work in terms of real fields when applying Poynting's theorem) and therefore time-averages to zero. The z-directed power <^> is given by — sech 2(/c1z0) + 1 + tanhOqzo)
(6.85)
We now have the complete solution in principle, since (6.81), (6.82) and (6.85) determine qx, A and z 0 in terms of a; and <^>. A full solution requires numerical work, since all three equations are nonlinear. However, a number of qualitative features can be extracted from the solutions as they stand. First, since tanh(/c xz0) < 1, (6.83) shows that a solution is possible only if K2 < K1, or equivalently s2 > e^ Thus for self guiding to occur, the linear 'substrate' 2 must have a refractive index higher than the low-power refractive index e{/2 of the nonlinear 'cladding' 1. Second, (6.82) shows that z 0 must be positive, since from (6.81) A is positive. It is then helpful to sketch the right-hand side of (6.85), as is done in Fig. 6.20. The function plotted, namely the expression in square brackets in (6.85), has a minimum value of 2. This makes it plausible that there is a self-guided solution only if <^> exceeds some threshold value, as we anticipated on general grounds. Fig. 6.20 Sketch of (6.85) versus K^Q.
217
Surface polaritons Results of numerical calculations for the self-guided s-polarised mode discussed here are shown in Fig. 6.21. This shows both the real and imaginary parts of qx, the latter being evaluated by a perturbation method by use of Poynting's theorem to find the power loss per unit path length. The existence of a threshold power level shows clearly on the figure. Various extensions of the calculation reviewed in this section have been carried out. Wendler (1986) generalises the calculation to include the presence of a surface-active film between the two semi-infinite media. For a nonlinear film sandwiched between linear media, (6.76) still holds in the film, but the constant of integration K need not be zero since there is no boundedness condition. The field E in the film can then be expressed in terms of the Jacobi elliptic functions (Boardman and Egan 1985), of which (6.79) is a special case. As stated, further details and references are given by Stegeman et al. (1986). 6.5 Attenuated total reflection 6.5.1 Basic principles A basic property of the surface polariton discussed in §6.2 is that it is confined to the region of the coqx plane satisfying q\ > where, as in Fig. 6.9, ex is the dielectric constant of medium 1, occupying the half-space z > 0. This means that the surface mode cannot be excited by a light beam incident in the normal way in medium 1, since such a beam Fig. 6.21 Real and imaginary parts qf and q\ for a wave guided along the interface between a linear and a nonlinear dielectric. Parameters are e t = 2.4025-0.00 li, OLl/celel = 10" 9 m 2 W~ 1 , and e2 = 2.4336 or 2.56. Also shown are the field distributions for various q* values for s2 = 2.4336 [after Stegeman et al. 1986].
1.70
1.60
-I
-
-
—
—-5.
- 0.4 x 10-
- 0.3 xlO"3
s - 0.2x10"
(
\ = 2.4336
-—
Pi
1
i
50
100
.
Pn i 150
Power (mW mm"1)
218
/»,
i 200
0.1x10
Attenuated total reflection with qz real. For this reason the surface polariton satisfies ql + q2 = s^/c2 is called 'non-radiative'. The restriction to real qz is removed in the method of attenuated total reflection, ATR for short. Here we give an introductory account; more detail is to be found in review articles by Otto (1974, 1976), by Borstel et al. (1974) and by Abeles (1986). Three ways of coupling an incident beam in medium 1 to surface polaritons are illustrated in Fig. 6.22. In the grating method, Fig. 6.22(a), a diffraction grating is laid down on the surface of interest. The incident light has surface wavevector qx = s{/2(co/c) sin 0,. By Bloch's theorem, (see §1.1), this couples to modes with qx values qx = s\/2(co/c) sin 6} + mln/d
(6.86)
where d is the grating periodic distance and m is an integer. Clearly for appropriate values of m and d coupling to nonradiative modes is possible. Fig. 6.22 Techniques for coupling light to surface polaritons: (a) grating method; (b) ATR, Otto configuration; and (c) ATR, Raether-Kretschmann configuration. In each case, the surface-active medium is shown shaded.
(a)
to 219
Surface polaritons When co, 9X and the order m are such that a surface mode is excited, the intensity of the diffracted light in that order is reduced. These reduced intensities observed with a diffraction grating are known as Wood's anomalies (Wood 1935). As a method of observing surface modes, the grating method has the disadvantage that the presence of the grating perturbs the properties of the mode under study. It is therefore not described further. Two different versions of the ATR method are sketched in Figs. 6.22(b) and (c). Both methods involve, strictly speaking, a three-layer geometry, and in both the light is incident in a prism of dielectric constant s1. In the Otto configuration, Fig. 6.22(b), a spacing medium, typically vacuum, is adjacent to the prism, with the surface-active medium as medium 3. The prism is chosen with a relatively large refractive index, ex > 62, and the angle of incidence 0, is greater than the critical angle 8C for total internal reflection, 01>8c = sin~1(el/2/s\/2). Thus if medium 3 were absent, the incident light would be totally reflected. However, total reflection requires the presence of an evanescent mode with decreasing amplitude (imaginary qz) in medium 2. The tail of the evanescent mode can excite a surface mode on the 2/3 interface; this surface mode removes energy, and thus produces an attenuation of the total reflection. Put a little more precisely, the in-plane wavevector qx = e\/2{co/c) sin 9X
(6.87)
is the same in all three media. Clearly with sufficiently large £ x and 6Y this can be made to satisfy the condition qx > e\/2(o/c
(6.88)
necessary for excitation of a surface mode at the 2/3 interface. Then when the incident frequency co and qx lie on the dispersion curve of a surface mode the reflectivity is reduced below unity. The main technical problem in the application of the Otto configuration is the control and uniformity of the spacer thickness, i.e. the thickness of medium 2. There is a short discussion of the experimental techniques in Otto (1976). In the Raether-Kretschmann configuration, Fig. 6.22(c), the surfaceactive medium is deposited direct on the face of the prism. Thus the method is most readily applicable when the material of medium 2 is easily evaporated, and the configuration has been mostly used to study surface plasmon-polaritons on metals such as Al and Ag. In this case s2 is negative, and the electromagnetic field is automatically evanescent in medium 2. The surface mode of interest is at the 2/3 interface, and the necessary 220
Attenuated total reflection condition for excitation is qx>sl/2co/c (6.89) in place of (6.88). As in the Otto configuration, the reflectivity is reduced when co and qx lie on the dispersion curve. Two forms of experimental scan are possible in ATR in either geometry. First, there is a frequency scan with angle 0, fixed. It follows from (6.87) that in this case a straight-line trajectory is scanned in the coqx plane. Second, there is a scan over angle atfixedfrequency; the trajectory is then a straight line parallel to the qx axis. The two possibilities are shown schematically in Fig. 6.23. In either case, where the scan trajectory crosses the surface-mode dispersion curve a dip should occur in the reflection coefficient, since the surface mode is excited. As previously remarked, both the Otto and the Raether-Kretschmann configuration involve a three-medium geometry. Thus the properties of the surface mode at the 2/3 interface are perturbed by the presence of the 1/2 interface. If the thickness D of medium 2 is too small, then the system is 'overcoupled', and the properties of the isolated 2/3 interface will not be observed. If D is too large, on the other hand, the system is 'undercoupled', and the 2/3 interface mode is only weakly excited, so that only a small reflectivity dip is seen. Choice of optimum D is therefore an important part of experimental design. Fortunately, as we shall now see, the theoretical expression for the reflectivity is readily derived, and numerical simulations can be performed to give a guide to the choice of D value. Fig. 6.23 Schematic comparison of frequency scan and angular scan in ATR [after Otto 1974].
Surface polariton
221
Surface polaritons 6.5.2 Theory The theory of ATR in either configuration is formally quite similar to the theory of the two-interface polaritons given in §6.3. Since surface polaritons are p-polarised modes with the H field in the y direction and the E field in the xz plane, ATR occurs only with p-polarised light, in which the E field is similarly in the xz plane. In medium 1, the E field takes the form E = (£ 0 , 0, qxE0/qlz) exp(iqxx - icot)
exp(-iqlzz)
+ r(£ 0 , 0, - qxE0/qlz) exp(iqxx - icot) Qxp(iqlzz)
(6.90)
Here the first term describes the incident wave, and the constant Eo is regarded as given. The second term is the reflected wave; the object of the calculation is to find the complex amplitude reflection coefficient r. In media 2 and 3, the £ field can be written in the form of (6.51) and (6.52). The next step in the calculation is to apply the two electromagnetic boundary conditions at each interface; this gives four equations for the four coefficients r, a, b and E3x in terms of Eo. Solution of these equations is straightforward, and yields r= z) + exp(-2iq2zL)(s3q2z z) + exp(-2iq2zL)(e3q2z
- e2q3z)(e2qlz - s2q3z)(e2qlz (6.91)
Equation (6.91) is a kind of response function, and like other response functions it is most useful when damping is included. However, one or two comments are possible about its significance in the absence of damping. The condition for the numerator to vanish is formally identical to (6.59), while the condition for the denominator to vanish gives (6.59) with the sign of qlz reversed. This is because (6.59) is the condition for Maxwell's equations to have a solution with only one field term in medium 1, namely (6.50), whereas (6.91) is derived from (6.90), in which two field terms are present. Experimental work to date has mainly concentrated on the intensity reflection coefficient R = |r| 2 , although as pointed out by Abeles (1986), for example, by using the techniques of ellipsometry it is possible to extract more information by measuring both the amplitude and the phase of r. Our discussion will largely be concerned with R. 6.5.3 Experimental results ATR curves for an Ag specimen in the Otto configuration are shown in Fig. 6.24. The importance of choice of gap width L is quite 222
Attenuated total reflection clear. For large L, curve a, the system is undercoupled, and the reflectivity dip is small. For optimum L, curve b, the reflectivity drops almost to zero. For small L, curve c, the system is overcoupled. The ATR dip is then broadened and its minimum is shifted from the point found in optimum coupling. The Otto configuration has more often been applied in the infrared, usually to study surface phonon-polaritons. Experimental results for GaP are shown in Fig. 6.25; analysis of these results gives very good agreement with the theoretical curve for the GaP/vacuum surface polariton, as is seen in Fig. 6.26. Experimental results for an angle scan on Au in the RaetherKretschmann geometry are shown in Fig. 6.27. As may readily be checked from the value quoted for the dielectric constant of the prism, the critical angle for total internal reflection at a prism/vacuum interface is 6C = 33.8°. For 8 > 6C the experiment would give R = 1 were it not for two mechanisms: the small loss due to the imaginary part of e2, and the reflectivity dip due to excitation of the surface plasmon at the Au/vacuum interface. The first of these shows up in the plateau region around 34°, and the second of course is responsible for the main dip. It is clear from these results that the ATR method is capable of precise determination of Fig. 6.24 Experimental and theoretical spectra for Otto configuration with glass prism, vacuum spacer and Ag specimen; detailed geometry in inset. Cross, experimental points; full line, theoretical curves obtained by least-squares fit of (6.91). The curves correspond to an angle scan at He-Ne wavelength k = 633 nm. Curve a, gap width L = 951 nm, undercoupled; curve b, gap width L = 918 nm, optimum coupling; and curve c, gap width L = 581 nm, overcoupled. [Courtesy of J. R. Sambles and K. Welford.] 1.40
silver 0.00
38.00
38.60
39.20 39.80 Angle (degrees)
40.40
41.00
223
Surface polaritons
both the real and imaginary parts of the dielectric function without recourse to a Kramers-Kronig analysis. In §6.2.5 it was mentioned that ATR was one of the principal techniques used for study of surface exciton-polaritons. Comparison of §6.4.2 with §6.2.5 shows how the theory must be adapted for this case. For the Otto Fig. 6.25 ATR spectra (frequency scan) for GaP/air interface with Si prism. The curves are labelled by values of qx taken from (6.87) where k0 = co/c. Gap values were L = 40, 25, 12, 2.5 fim, increasing as qx increases [after Marschall and Fischer 1972].
f
1 (JL
100 -
N 1.023 A:0
qx =
V A i(l ;
/
\
80 -
-
1.056& 0
1 125 A:
Al
0,
2.00 k0
A
-
60 360
380 a)/2nc (cm"1)
400
Fig. 6.26 Dispersion curve of GaP/air interface. Theoretical (full line) and experimental (squares). GaP parameters as in Fig. 6.25 [after Marschall and Fischer 1972]. 0)
I
/
x=
0J C
/
400 I 1 1 / 1 / 1 I
(i)/2nc (cm"1)
l
380
360
/
i /
I
/
I
1.0
224
i
i
i
1.5
2.0
2.5
cqjo)r
Attenuated total reflection geometry, the surface-active medium is medium 3, and as in §6.2.5 it is described by three field amplitudes (two transverse, one longitudinal) rather than the one used in the derivation of (6.91). Corresponding to the two additional amplitudes, there are two extra boundary conditions, derived for example from (6.13). Thus it is still possible to find an expression for the reflected amplitude r, although the expression is more complicated than (6.91). The first calculation of this kind was reported by Maradudin and Mills (1973). It was pointed out in §6.5.1 that for the Raether-Kretschmann geometry to be used the specimen must be evaporated on to the coupling prism. Thus this geometry is not applicable for excitonic materials. The use of the Otto configuration is difficult, however, because the thickness of the space layer must be comparable to an optical wavelength, which now means typically a few hundred nanometres. Some discussion of experimental techniques is given by Tokura et al. (1981). Fig. 6.28 shows results obtained by these authors on ZnSe specimens with an evaporated MgF 2 spacer layer. It would take us too far to go into the details of these results or the discrepancy with the theoretical curves. It may be noted, however, that the authors divide their specimens into Type A and Type B, the former having freshly cleaved surfaces while the latter have surfaces that have been Fig. 6.27 ATR in Raether-Kretschmann geometry: angle scan on gold specimen at wavelength X = 700 nm. The theoretical curve is a least-squares fit of (6.91) to the data,yielding £i = 3.2217,e 2 = -16.781 + 1.3169i and L = 45.355 nm.ej and L are included in the fitting parameters since it can then be checked that the same values are obtained from fits to data obtained at different wavelengths. [Courtesy of R. A. Innes and J. R. Sambles.] 1.20
1.00
•t °-80 ^ 0.60-1 0.40 0.20 0.00
33.60
34.06
34.52 34.98 Angle (degrees)
35.44
35.90
225
Surface polaritons
exposed. The differences between the two types of spectra bring out the sensitivity of excitonic effects to surface conditions. Both bulk and surface exciton-polaritons are low-temperature excitations, since the exciton becomes ionised at moderate temperature. This is illustrated in Fig. 6.29, which shows the disappearance of Type B spectra with increasing temperature. As depicted in Fig. 6.22, a diffraction grating can be used to couple light into a nonradiative mode such as a surface plasmon, and equally a grating can couple light out. In a similar way, a rough surface couples light out. A rough surface can be regarded as a superposition of gratings Fig. 6.28 ATR spectra in p-polarisation for ZnSe in Otto geometry at 2 K. Gap distances 100 nm (Type A) and 90nm (Type B): critical angle of incidence 55°. The curves show frequency scans, with angles of incidence marked. Broken curves are theoretical, using form of ABC proposed by Rimbey and Mahan (1974) [after Tokura et al. 1981].
2.80
2 81
2.80
Energy (eV)
226
2.81
Attenuated total reflection of different periodicity, or equivalently one can say that a rough surface destroys translational invariance in the surface plane. Either way, it is clear that light can be emitted. The surface of an Ag film evaporated on quartz is already rough enough to produce significant coupling, and indeed light emission was observed in the earliest experiments (Kretschmann and Raether 1968). It was this radiative property of a rough surface that was exploited in the nonlinear-generation experiment of De Martini et al. (1977), sketched in Fig. 6.15(a). Much experimental and theoretical work has been carried out on rough surfaces; the reader may refer to the reviews by Raether (1982) and by Maradudin (1982) and to the more recent theoretical review by Maradudin (1986). In this section we have described the principle of the ATR method and presented experimental results obtained on three-medium geometries as illustrated in Figs. 6.22(ft) and (c) and analysed in (6.91). In fact the method has been applied widely to study more complicated specimens consisting of a number of layers. As an example, Fig. 6.30 shows an angle-scan ATR curve for a prism/MgF2/Ag/air geometry, that is, the Otto configuration with MgF2 as the spacer and Ag as the surface-active medium. Although the Ag film is bounded unsymmetrically, with MgF2 on one side and air on the other, nevertheless the higher-frequency mode behaves like the long-range mode (LRSP) that was described for a Fig. 6.29 Temperature dependence of Type B spectrum of Fig. 6.28 [after Tokura et al. 1981]. TypeB 6 = 65.4°
2.76
2.78 2.80 Energy (eV)
2.82
227
Surface polaritons
symmetric geometry in §6.3. This is clearly seen in Fig. 6.30, which shows that the high-frequency mode (lower angle) is a much sharper resonance than the low-frequency mode. ATR has also been extensively applied to the investigation of adsorbed molecular layers on Ag; a review is given by Abeles (1986). 6.6 Other experimental methods 6.6.1 Raman scattering
As mentioned at the end of §6.1.1, much of the experimental investigation of bulk polaritons was carried out by means of Raman spectroscopy. By contrast, compared with ATR spectroscopy, Raman scattering has yielded only a relatively small amount of information concerning surface polaritons. The combination of a small scattering volume with the need to employ near-forward scattering angles makes the experiments technically demanding. In fact the only results published to date are by Ushioda and co-workers. Some of their key results are presented in this section; a fuller account is given by Ushioda and Loudon (1982). It was thought at one time that surface polaritons might be seen in backscattering from the surface of a semi-infinite medium. This is true in principle, but the cross section is expected to be smaller by a factor of 102 or 103 than that for forward scattering from a slab specimen and, Fig. 6.30 ATR curve (angle scan) for a system comprising prism/MgF2/Ag/air at a wavelength of 632.8 nm. Cross, experimental points; full line, least-squares fit of theory, giving thicknesses 203 and 28.2 nm for MgF 2 and Ag films respectively. [Courtesy of J. R. Sambles and M. D. Tillin.] 1.20
0.00
228
48.2 55.! Angle (degrees)
63.4
71.0
Other experimental methods in fact, backscattering has never been observed. The reason was first explained in detail by Chen et al. (1975). The theory is analogous to that given for Brillouin scattering in §2.5, and in particular the backscattered field contains the factor (Qz + k\2 + kl2)~l seen in (2.116). Here fcf2,fc|2are the z components of the wavevectors of the incident and scattered light. It is seen that they add to generate a large denominator for backscattering. By contrast, for near-forward scattering the corresponding factor is (Qz + k\2 — k^)" 1 ; the z components of the optical wavevectors now subtract to produce a small denominator and consequently a larger cross section. Experimental results for the single-interface polariton were reported by Valdez and Ushioda (1977). The experiments were carried out on a GaP specimen of thickness L = 20 /mi. Correct choice of L was crucial. The argon-ion laser frequency is close to the GaP band gap, so that resonant enhancement of the cross section occurs (Hayes and Loudon 1978) and the specimen is fairly opaque; L is chosen large enough so that the upper- and lower-branch polariton curves effectively coincide on the single-interface curve, and small enough for some light to be transmitted. Results for the scattering intensity are shown in Fig. 6.31. The curves are Fig. 6.31 Raman spectra in a single-crystal slab of GaP with (111) faces. L= 20/mi. The curves are labelled by scattering angle 6 defined in inset [after Ushioda and Loudon 1982].
•a
e = 2.o° 6= 1.6° 390
395
400
405
410 1
Frequency shift (cm" )
229
Surface polaritons dominated by the large peak due to the LO phonon at a)hO/2nc = 403 cm" 1 . The surface-polariton scattering is the small feature marked by an arrow. It is seen that the frequency of this feature changes with #, corresponding to the dispersion of the surface-polariton curve. The peak position is plotted and compared with the theoretical dispersion curve in Fig. 6.32. The slight discrepancy between experiment and theory is thought to be due to surface roughness. The first experimental results to be published were in fact for the lower of the two modes of GaAs on sapphire in a two-interface geometry (Evans et al. 1973). In that experiment, the upper mode was not resolvable from the large peak due to scattering off the bulk LO phonon. Subsequently the technique was refined by Prieur and Ushioda (1975) so that the upper-mode peak could be detected. They made two essential changes from the earlier experiment. First instead of air/GaAs/sapphire they used benzene/GaAs/sapphire; the higher refractive index of benzene means that the upper-mode frequency is lowered away from the bulk LO frequency. Second, Prieur and Ushioda made use of the fact that the surface modes are not detected in backscattering, whereas the bulk-mode peaks are identical in forward and backward scattering. They therefore devised a subtraction technique to measure the difference between the two, which contains only surface-mode scattering. Scattering data are illustrated in Fig. 6.33, and the resulting dispersion curves are shown in Fig. 6.34. The possibility of observing guided-wave polaritons emerged from theoretical work by Mills et al. (1976) and by Subbaswamy and Mills (1978), who predicted that the scattering peaks should be comparable in intensity to those from surface-type modes, and therefore observable. Fig. 6.32 Surface-polariton peak positions of Fig. 6.31. Theoretical curve for air/GaP interface using dielectric function of (5.60) with coT/2nc = 367.3 cm" 1 , cojlnc = 403.0 cm" 1 and ^ =9.091 [after Ushioda and Loudon 1982]. 400 r-
r 398 396
394 2
3
Scattering angle (degrees)
230
Other experimental methods Fig. 6.33 Extraction of surface-mode Raman scattering peaks by the subtraction technique of Prieur and Ushioda (1975) [after Ushioda and Loudon 1982].
c
Frequency (cm l)
Fig. 6.34 Dispersion curves for upper and lower surface modes in Benzene/GaAs/Al 2 O 3 . Theoretical curves correspond to (6.69) [after Ushioda and Loudon 1982].
290
L
285
280
275
270
231
Surface polaritons Scattering data obtained on a 30 /mi single-crystal GaP slab are shown in Fig. 6.35, with the dispersion curves shown in Fig. 6.36. The publication of the first experimental results stimulated considerable theoretical effort. As was seen in the example treated in detail in §2.5.2, there are several steps in the development of a full theory that will account for intensities and lineshapes as well as peak positions. It is first necessary to find the equivalent of (2.111) describing coupling of the incident light amplitude E, at frequency cox to the field at frequency co to produce a polarisaton P s at frequency cos = co1 ±CO. AS with (2.111), the coupling mechanism is identical to that for scattering by the corresponding bulk modes. The polariton is a coupled mode containing both an electric field amplitude E and an optic phonon amplitude W. The polarisation P s contains contributions from both amplitudes: PI = a*fiEf2W* + bapyE?2Ey* (6.92) where summation over repeated indices is implied. The coupling tensors a and b are discussed by Hayes and Loudon (1978). Fig. 635 Near-forward Raman scattering by guided-wave polaritons in GaP slab. Curves are labelled by scattering angles [after Ushioda and Loudon 1982]. GaP 30fim
250
232
(111)
300 350 Frequency (cm"1)
400
Other experimental methods
The calculations for the transmission of the incident light into the specimen and the radiation of scattered light by the polarisation P s are identical to those outlined in §§2.5 and 2.6. The cross section is proportional to the power spectrum of the scattered field, and therefore is ultimately determined by power spectra involving the fields W and E of the polariton. These are conveniently evaluated from Green functions such as <<£a(Q); ^ ( Q ' ) ) ) ^ by means of the fluctuation-dissipation theorem. Thus, as in all light-scattering calculations, the underlying problem can be expressed as the evaluation of appropriate Green functions of the scattering mode. Calculations along the lines indicated here were first presented in full by Nkoma and Loudon (1975), Nkoma (1975) and Mills et al. (1976); these and subsequent papers are reviewed by Cottam and Maradudin (1984). 6.6.2 Light-emitting tunnel junctions
A method of excitation of surface plasmons that has attracted attention is to use the tunnel current flowing between two metals. Under Fig. 636 Dispersion curves of guided-wave polaritons in GaP slab. Experimental points from Fig. 6.35 and similar data; theoretical curves from (6.64) and (6.65) [after Ushioda and Loudon 1982]. 0= 1.0° 1.3°
1.8° 2.0°
3.0°
233
Surface polaritons
some circumstances, for example if the interfaces are rough or if the substrate is a diffraction grating, the junction is seen to emit light, typically in the visible part of the spectrum. An example of light emission by a tunnel junction is shown in Fig. 6.37, with the specimen configuration shown schematically in the insert. The first layer deposited on the glass substrate is CaF 2 ; this has a rough top surface, and consequently all subsequent interfaces are rough. The tunnel junction itself is Al/Al2O3/Ag, the A12O3 tunnelling barrier being formed by oxidation of the Al film. In the experiment a dc bias voltage Vo is applied between the Al and the Ag, and a tunnel current / 0 flows through the A12O3 barrier. The spectrum of light emitted by the junction is seen to change from red to blue as Vo increases. The actual spectrum emitted in the direction of the specimen normal is shown in Fig. 6.37. It should be noted that the spectrum is unpolarised, and the maximum photon energy hcoQ, corresponding to the short-wavelength cut off of the spectrum, is given to a good approximation by ha>0 = eV0. Thus at the short-wavelength end virtually all of the energy of a tunnelling electron is converted into photon energy. Two other important forms of tunnel junction are first a junction deposited on a diffraction grating, and second a junction in which the top electrode consists of discrete metal particles. In the first case the spectrum is p-polarised and has a sharp peak whose position varies with angle of Fig. 6.37 Spectra of light emitted by rough Al/Al 2 O 3 /Ag tunnel junction in normal direction. Curves are labelled by tunnelling current / 0 . Inset shows specimen geometry [after Dawson et al. 1984]. Thickness (nm) 25-40 2-3 f50-70 100-140
125
100
•s
75
50
25
400
500
600 A (nm)
234
700
Surface magnon-polaritons emission relative to the specimen normal. In the second case the spectrum is predominantly p-polarised but broadband. In all three types of light-emitting tunnel junction the tunnel current excites a surface plasmon which then decays radiatively because of the roughness to produce the observed spectrum. The nature of the plasmon excited depends on the form of the junction; for a detailed discussion and further references the reader may turn to Dawson et al. (1984). 6.7 Surface magnon-polaritons In this final section we turn to polariton effects arising from frequency dispersion in the magnetic susceptibility. These have already been discussed in Chapter 4 within the magnetostatic approximation; this merited a separate chapter because of the extensive experimental results that have been obtained, particularly by light scattering. We now deal with the longer wavelength region where inclusion of retardation is important; that is, we are concerned with the electromagnetic region of Fig. 4.1. The results to be presented are mainly theoretical, since very few experimental results have been reported to date. However, there is reason to believe that more experimental results will soon be available. Our discussion is largely based on that in earlier reviews by Sarmento and Tilley (1982) and by Tilley (1986). As in Chapter 4, we consider a magnetic material (ferromagnet or antiferromagnet) with a static magnetic field Bo directed along the z axis. The magnetic susceptibility tensor in the xy plane takes the gyromagnetic form of (4.15) where the form of xa a n d Xb depends on the material under discussion. As demonstrated in Problem 4.1, the susceptibility is diagonalised by transformation to the rotating-wave representation, (6 93) m±=mx±imy h±=hx±ihy (6.94) + ± ± ± with susceptibilities x > X defined by m =x h . Expressions for the susceptibility components of a ferromagnet are given in (4.16), (4.17) and Problem 4.1; for an antiferromagnet the corresponding expressions are (4.54) to (4.56). 6.7.1 Bulk-magnon polaritons The bulk polariton modes were first discussed by Auld (1960) for a ferromagnet; later authors dealt with antiferromagnets (Manohar and Venkataraman 1972; Bose et al. 1975; Sarmento and Tilley 1977) and ferrimagnets (Oliveira et al. 1979). Maxwell's equations lead to q2h - q(q-h) = (sa)2/c2)(h + m)
(6.95) 235
Surface polaritons for a plane wave exp(iq-r — icot). Since the static magnetic field is in the z direction, there is no loss of generality in choosing qy = 0. Equation (6.95) then yields h+ =£1m+
-£2m+
(6.96)
JT = -f 2 m +<*!?*-
(6.97)
^ = ±(2eco2/c2 - ql)l{q2 - eco2/c2)
(6.98)
where 2
2
2
2
Z2 = h J(q -™ /c ) When combined with the susceptibility equations m±=x±h±
(6.99) (6.100)
(6.96) and (6.97) yield the polariton dispersion relation (i-Za+)(i-Za-) = Z22X+x(6.101) which is a general form. It is clear from (6.98) and (6.99) that, as might be expected on general grounds, the dispersion relation depends on the angle 6 between the propagation vector q and the z axis. Dispersion curves for a ferromagnet are shown in Fig. 6.38. It is seen that one solution of (6.101) shows photonmagnon mixing, while the other does not. The results presented so far are concerned with the solution of homogeneous equations of motion for the magnetic system coupled to the electromagnetic field. A fuller treatment involves finding the Green functions as the linear response to an applied rf field hext exp(iq»r — icot). This has been done for bulk ferromagnets (Sarmento and Tilley 1976) and antiferromagnets (Sarmento and Tilley 1977). 6.7.2 Surface magnon-polaritons: ferromagnets We now turn to the surface polaritons at the interface between vacuum and a magnetic medium. We concentrate on the Voigt configuration, in which B o is in the plane of the surface and propagation is perpendicular to B o , since most of the results that have been derived concern this geometry. As in Chapter 4, we take the z axis along B o and the x axis as the normal to the surface, so that the propagation vector qLi is in the ±y direction. The magnetic medium is in the half-space x < 0. Following Hartstein et al. (1973), who were the first to investigate this problem, we work in terms of a gyromagnetic permeability tensor (6.102)
236
Surface magnon-polaritons where = Ml + la)
(6.103) (6.104)
Vxy =
(6.105) This is derived as in §4.2 with the addition of a constant 'background permeability'. Such a term can only arise from the contribution of resonances at a higher frequency than those appearing explicitly as poles of X- It is therefore plausible, for example, for a ferrimagnet at microwave frequency where )U could be due to the exchange resonance. It is worth retaining \i explicitly, since as will be seen its presence leads to the appearance of a mode that is absent for JX = 1. For the magnetic medium, Maxwell's equations now give V2h - V(V-h) - (e/c2)fi d2h/dt2 = 0
(6.106)
and 0
(6.107)
Fig. 6.38 Dispersion curves for bulk polaritons propagating along the static magnetic field (0 = 0) and perpendicular to it (0 = 90°) in a bulk ferromagnet. Bo = fi0M0 = 0.175 T [after Auld I960].
237
Surface polaritons while in the vacuum V2h-(l/c2)d2h/dt2
=0
(6.108)
In order to find the dispersion relation of the surface mode, we substitute into (6.106) to (6.108) and the boundary conditions the ansdtze h = h2 exp(iq || y) exp(/c2x) exp( - icot) h = h : exp(ign j;) exp( — KXX) exp( — icot)
x<0 x>0
(6.109) (6.110)
Substitution of (6.109) into the z component of (6.106) gives h2z = 0, and the boundary condition on tangential h gives hlz = 0. It follows that the surface polariton is s-polarised with E in the z direction. The other two components of (6.106), with (6.109), yield K2-ql^(sco2/c2)fiy
=0
(6.111)
where fiy is called the Voigt permeability Vv = l*xx + t*ly/Hxx Equations (6.108) and (6.110) give K2-ql+CD2/c2 = 0
(6.112) (6.113)
Equations (6.111) and (6.113) are the analogues of (6.15) for the dielectric problem; they characterise the individual media with no reference to boundary conditions. The Voigt permeability is central to this geometry. For the particular case of a ferromagnet, substitution of (4.16) and (4.17) gives t**/l* = C K + ^m) 2 - ^ 2 ] / K +
+
'iq^Jii^
(6.115)
As is seen from (6.104), as long as the resonance modes are undamped jj,xy is pure imaginary, and (6.115) involves purely real quantities. Equations (6.111), (6.113) and (6.115) are simultaneous equations for the three quantities K19 K2 and q^. Elimination of K1 and K2 yields an equation, quoted explicitly by Hartstein et al. (1973), for the dispersion function q\\((o). The most important property of the dispersion curve can be seen from (6.115). Since q^ occurs there linearly, the directions -\-q^ and —q y are not equivalent. Thus, as with the magnetostatic modes of Chapter 4, the propagation is nonreciprocal. 238
Surface magnon-polaritons The dispersion curve for a ferromagnet found by numerical solution of (6.111), (6.113) and (6.115) is shown in Fig. 6.39. The +ql{ mode is the generalisation of the Damon-Eshbach mode to include retardation. For large q^ the curve coincides with the Damon-Eshbach one, being asymptotic to the surface frequency cos of (4.31), which with \x included becomes <*>s = l>o + Kco0 + comy]/(n + 1 ) (6.116) For small q^ the Damon-Eshbach curve has the unphysical property that it crosses the vacuum photon line and reaches q^ = 0 at the frequency a>B of (4.28). The calculation with retardation corrects this; the mode still terminates at coB, but at the nonzero value qy given by
239
Surface polaritons To date there has not been any experimental observation of surface ferromagnetic polaritons. The frequencies are around co0 ~ 1 GHz, with wavenumbers of order CDO/C. In light scattering, therefore, the scattering angle would have to be impracticably small to get down from the magnetostatic into the polariton region. Matsuura et al. (1983) suggested that ATR might be applied, but since the wavelengths are of order 1 cm or greater, specimens would have to be very large.
6.7.3 Surface magnon-polaritons: antiferromagnets The difficulties of observation of surface magnetic polaritons are in principle reduced for antiferromagnets, since resonance frequencies are then in the far infrared and specimens need not be so large. For a uniaxial ferromagnet the results of the previous section apply as long as the appropriate expressions (4.54) and (4.55), are used for xa a n d Xb m (6.103) and (6.104). The properties of surface polaritons on an antiferromagnet were discussed by Camley and Mills (1982). One important result is that, as for the magnetostatic mode discussed in §4.5, in zero applied field the propagation is reciprocal, cos(q^) = cos( — ql{). In the presence of an applied field, however, the propagation is nonreciprocal. In addition to the dispersion curve, Camley and Mills present calculations and numerical results for ATR involving surface antiferromagnetic polaritons. They conclude that ATR is viable, but to date no experimental results have been reported. Optical reflection measurements on MnF 2 by Remer et al. (1986) give indirect evidence of the nonreciprocity of surface polariton propagation in the presence of an applied field. To understand how this comes about one may argue loosely as follows. Ordinary reflectivity probes the region to the left of the light line, whereas the surface-mode dispersion curve lies to the right of the light line. In the presence of damping, however, the dispersion curve becomes broadened and the properties of the surface mode are echoed, to some extent, in the reflectivity. Remer et al. substantiate this argument by calculating the reflectivity with a damping time T included in the susceptibilities xa a n d XbMnF 2 has a relatively low anisotropy field and therefore a low resonance frequency. For this reason, Remer et al. were able to use a microwave source at 88 GHz driving a third-harmonic resonator to produce a fixed-frequency source at 264 GHz. The results are presented as reflectivity versus magnetic field. Figure 6.40 shows calculated and measured reflectivity curves for two different directions of applied field B in the plane of the surface. The results clearly show the expected nonreciprocity. 240
Surface mag non-polar itons 6.7.4 Two-interface modes The magnetic polariton modes of a film of finite thickness L in vacuum were first investigated by Karsono and Tilley (1978) and by Marchand and Caille (1980). We use essentially the notation of Fig. 6.17, but with x and z axes interchanged. Analogously to (6.50) to (6.52), the magnetic field is taken in the form 11 = 11! exp(i^f|| y) exp( — icot) exp( —K XX)
X>0
(6.119)
H = exp(i || y) exp( - icuf)[A exp(iqxx) + B exp( - iqxx)~] 0>x>-L H = H 3 exp(i^ny)exp( — icot) exp(/c3x)
x < —L
(6.120) (6.121)
where qx is real for guided modes and imaginary, qx = iK2
(6.122)
for surface modes. The wave equations and divergence equation, (6.106) to (6.108), continue to apply, and in consequence (6.111) for K2 (or qx) and (6.113) for KX are still valid. The dispersion equation is found by applying standard boundary conditions, and the result for e = 1 as quoted by Marchand and Caille (1980) is (KJri + K22 + q\\ilyl\i2xx) tanh(/c 2 L) + 2K 1 IC 2 JI V = 0
(6.123)
This is the generalisation of the Damon-Eshbach result for a ferromagnetic film, (4.27), to include retardation as well as the extension of the result of Hartstein et al. (1973) to a finite thickness slab. Fig. 640 Reflectivity at 264 GHz off the surface of MnF 2 at 28 K. (a) Calculated with G>AF= 1.24 THz and (COAF^) 1 =6.5 x 10~ 4 , where O>AF is zero-field resonance
frequency, (3.86). Damping time x is chosen to give best fit to experimental results. {b) Measured [after Remer et al. 1986].
R (arb. units)
R (arb. units) . . . B~ — B+
(a) 0.8 0.6 0.4 0.2
o B-
(«
1 MJ
A
J i
0.6
ii
y— -
1/
0.8
B(T)
Ik iii \\M
ill
i
0.4
I
I
0.6
I
0.8 B(T)
241
Surface polaritons The surface mode solutions are those for which K2 is real and qx imaginary. Dispersion curves for the +q^ mode, calculated numerically, are shown in Fig. 6.41. Like the magnetostatic mode and the surface mode on a semi-infinite medium, the mode starts at the frequency coB of (4.28) and is asymptotic for large q^ to the surface frequency cos of (6.116). For the larger value QB = 2 of applied field, Fig. 6.41 (a), there is considerable separation between the dispersion curves and the semi-infinite result even for \y\MoL/c = 0A5. Since the scaling distance c/|y|M 0 is quite large, the finite-thickness corrections are likely to be of importance. The dispersion
Fig. 6.41 Dispersion curves for a ferromagnet with n= 1.25. Full line, + qj retarded slab mode from (6.123); broken line, + | mode of Fig. 6.39; chain line, magnetostatic modes, (a) Q B = #0/^0^0 = 2; (b) Q B = 0.005. The curves are marked by values of \y\M0L/c [after Karsono and Tilley 1978]. (a)
2.56
2.50
a
2.44
^ 3 II
10
15
1.0
1.5
0.5
a
0.5
242
Surface magnon-polaritons
curves are well separated from the magnetostatic modes for QB = 2 but rather close for QB = 0.005; thus the effects of retardation show up more in strong applied fields. As with the dielectric guided-wave polaritons of Fig. 6.19, there are two frequency 'windows' for the guided modes. Typical dispersion curves in these windows are shown in Fig. 6.42. The lower-frequency modes start Fig. 6.42 Guided-mode dispersion curves (full line), bulk magnetic polariton (chain line) and vacuum light line (dotted line), (a) Lower frequency window, ^t = 1.75, QB = B0/n0M0 — 2, \y\M0L/c=l. {b) Upper frequency window, /x=1.75, Q B = 10, \y\M0L/c = 0.5 [after Marchand and Caille 1980]. (a)
18
12 II
10
20
15
(b)
31
3 II
a 21
21
31
41
cqt/\y\M0
243
Surface polaritons
on the light line K1 = 0 and have the short-wavelength asymptote Q -» Qv. In the high-frequency window, the lowest mode starts on the bulk magnetic polariton curve, given by qx = 0, and the higher modes again start on the light line. At high frequencies all the modes are asymptotic to the bulk polariton curve. As for the surface polaritons on a semi-infinite specimen, there have as yet been no experimental studies of these retarded slab modes. Fukui et al. (1984) have calculated ATR spectra for the Karsono-Tilley type modes, but just as with a single-surface mode the long wavelength implies a specimen of very large surface area. The outlook is presumably more hopeful for antiferromagnets, but to date no theoretical work has been published on dispersion curves or ATR spectra for polariton modes in an antiferromagnetic film. Just as for the bulk magnetic polaritons, more information can be obtained on the surface modes if Green functions are evaluated. This poses very considerable algebraic problems, but the polariton Green functions for a ferromagnetic film and semi-infinite medium have in fact been derived by Lima and Oliveira (1986). 6.8 Problems 6.1 The semiconductor G a ^ ^ ^ A s may be viewed as derived from GaAs by substitution at random Ga sites of a fraction x of Al. It exhibits two TO phonon resonances, one near to that of GaAs and one near to that of AlAs. The expression co\ — o r2 — I gives a good fit to IR reflectivity data (Kim and Spitzer 1979). For x = 0.3, numerical values are e^ = 10.16, col/2nc = 265.2 cm" *, TJlnc = 7.2 cm" \St = 1.24,co2/2nc = 360.6 cm" \ V2/2nc = 10.8 cm" \ S 2 = 0.93. (a) Sketch the real and imaginary parts of s(co) as functions of co. (b) Sketch the bulk-polariton dispersion curve. (c) Sketch the surface-polariton dispersion curve for a vacuum interface. (d) Write computer programs to draw the curves of (a) to (c) accurately. 6.2 The bulk exciton-polariton dispersion curve is described by (6.12) and (6.11) (with F = 0). Confirm the statement made in the text that there is one solution for co < coL and two for co > coL, and find the expression for coL. 6.3 Sketch and/or draw accurately the surface-polariton dispersion curve for an interface between vacuum and n-InSb in zero magnetic field, with an isotropic dielectric function given by (5.65).
244
Problems 6.4 Equation (6.42) gives a dispersion equation for surface polaritons on an anisotropic medium. When the (x, y, z) and (x', / , z') axes coincide it reduces to CO2 £ 1 £ z (£ x — £ j )
2
c2
sxez - s2
Derive this result from Maxwell's equations and boundary conditions. 6.5 In §6.2.4 the dispersion relation (6.46) for surface polaritons guided by a charge sheet was derived. Now suppose the charge sheet is replaced by a layer that can be polarised in the plane, so that (6.45) is replaced by (This might be realised by a suitable organic layer.) Prove that (6.46) is replaced by Kt
K2
Discuss this for the particular case = Sco2/(co2-co2) (a) Prove that Nakayama's result, (6.46), is the formal limit as col -• 0 with Scol = constant. (b) Sketch the dispersion curve. X(co)
6.6 Equations (6.62) and (6.63) give the dispersion relations for the surface modes of a symmetric slab. (a) Determine in which of the modes the electric-field pattern is symmetric about the mid-point of the slab, and in which it is antisymmetric. (b) Find the dispersion equations of the corresponding electrostatic modes first by letting c -• oo in (6.62) and (6.63), and second by generalising the potential analysis given in §6.2.2. 6.7 This is concerned with the calculation of the dispersion equation for the s-polarised guided modes. In place of (6.50) to (6.52) the E fields, which are all in the y direction, are written E — Ex exp{iqxx — icot) exp(ig
lz z)
z> 0
E = \c exp iq2 Jz + - J + d exp -iq2z(z + -J j x exp(igxx — icot)
0> z> — L
£ = E3 exp[ig 3z(z + L)] exp(igxx — icot)
z < —L
The boundary conditions involve the tangential magnetic-field component Hx, which is derived from the Maxwell equation for V x E. Show that the boundary conditions produce four homogeneous equations in the 245
Surface polaritons four amplitudes Et, c, d, E3:
cf-1+df=E3 where / is defined in (6.58). Show that these equations may be derived from (6.54) to (6.57) by the substitutions (6.66) and (6.67). 6.8 The number of solutions to (6.69) and (6.70) for s-polarised guided modes of a film depends on the value of L. Consider the case when zx and e2 are both real and positive, with £2 > £ i- Write the equations s2(o2/c2-ql)lf2L]
(6.69)
and
S^/f
/C
2 ?1
= -cot[i(e2co2/c2 - q2xyi2L]
(6.70)
(e2w2/c2-q2x)112 Sketch the left- and right-hand sides of these equations versus qx for various L values with co fixed. Hence show that (a) Equation (6.69) has one solution for small L, and the number of solutions increases as L increases. (b) Equation (6.70) has no solutions for small L. Find the value of L c such that for L< Lc the film is single-moded, i.e there is one solution of (6.69) and none of (6.70). Use (6.64) and (6.65) to investigate the corresponding results for p-polarised guided modes. 6.9 Check the derivation of (6.85) for the time-averaged power in the self-guided mode. 6.10 Confirm (6.91) for the ATR reflection amplitude.
246
7 Layered structures and superlattices
A number of crystal growth techniques have been developed for the production of specimens which have the form of a succession of layers. In molecular-beam epitaxy (MBE) beams of atomic or molecular species passing through ultra-high vacuum impinge on a single-crystal substrate and in the right conditions crystal growth occurs epitaxially, that is, with the crystal structures in register. In metallo-organic chemical vapour-phase deposition (MOCVD) growth occurs by deposition from a flowing vapour. Both MBE and MOCVD are used to produce semiconductor specimens in which, for example, single-crystal layers of GaAs and Al^Ga^^As alternate. Metallic specimens, including those in which one or both constituents may be magnetic, are made by sputtering. A review of MBE is given by Joyce (1985). Parker (1985) and Chang and Ploog (1985) include detailed chapters on MBE as well as on the physics of the resulting specimens, while the latter also contains a chapter on MOCVD. The growth techniques can be used to prepare specimens consisting of alternating layers of thickness dl of constituent 1 and thickness d2 of constituent 2. Specimens can be prepared so that dx and d2 have any value from two or three atomic spacings up to the order of 100 nm typically. We refer to them as periodically layered structures or superlattices;
the latter has a more specific meaning for electronic states in semiconductor structures, as we shall see, but it is often convenient to use it as a synonym for the former. Many of the physical properties are greatly modified by the existence of the long spatial period D = dl + d2. As mentioned already in §1.1.2, the most important general consequence is that a new Brillouin zone edge appears with wavevector component n/D perpendicular to the interfaces. This can be much smaller than the zone-edge wavevector n/a related to the lattice constant a. Dispersion curves, such as for acoustic phonons for example, develop band gaps at these new zone edges. 247
Layered structures and superlattices
Within the context of this book, it is natural to distinguish between the 'bulk' modes of an infinitely extended superlattice, and 'surface' or 'guided-wave' modes that may occur in specimens with one or two free surfaces. Since a 'bulk' specimen consists of an array of interfaces, an understanding of bulk modes uses many of the concepts developed in earlier chapters, and much of this chapter is therefore concerned with the bulk modes. As before, the emphasis will be placed on acoustic, optical and magnetic properties, although introductions to single-electron states and plasmons are given in §§7.4 and 7.5 respectively. The reader should be cautioned that up to now most of the interest in semiconductor superlattices has centred on the electronic properties because of the potential for device manufacture. The balance of topics is therefore consistent with the rest of the book but not representative of the literature. In §7.1 we introduce the main ideas with a detailed study of the continuum-acoustic modes of a bulk superlattice. Using a formalism very similar to that of Albuquerque et al. (1986a), we show that the system can be described by means of a transfer matrix T which relates the field amplitude at coordinate z + D to those at z, where z is measured perpendicular to the interfaces. The dispersion relation is simply expressed in terms of T. Subseqent sections are also largely based on the transfer-matrix formalism. In §7.2 we turn to the lattice dynamics of a simple ID model of a diatomic superlattice; despite its simplicity, this model has been widely employed in the analysis of experimental results. An introduction to optical properties, with emphasis on the reststrahl region, is given in §7.3. After a brief account of single-electron states in §7.4, we turn to collective-electron properties, that is, plasma effects, in §7.5, and finally §7.6 is an introduction to magnetic superlattices. As indicated, this chapter is concerned with semiconductor and metallic superlattices in which the long period is produced by variation of crystalgrowth conditions. Long periods also exist in some other types of material. Intercalated compounds, for example intercalated graphite, are basically layered materials in which long organic molecules are interposed to space out the layers of the starting material (see e.g. Dresselhaus 1986). Here the length of the organic molecules can be chosen to give a stipulated layer-to-layer spacing. Liquid crystals (see e.g. de Gennes 1974) containing chiral, that is screw-like, molecules can undergo a helical distortion in which the mean orientation of the molecules spirals round an axis in space; the two phases of most importance are the cholesteric and the smectic-C*. In these the pitch of the helix is a function of temperature and concentration of chiral molecules and can be varied in addition by 248
Continuum acoustics: folded acoustic modes
applied electric and magnetic fields; it can have any value from a few tens to many hundreds of nanometres. Clearly the properties of these long-period materials show analogies with those of superlattices. However, the analogies have not yet all been worked out, and it would take us too far afield to discuss these here. 7.1 Continuum acoustics: folded acoustic modes 7.1.1 Propagation in one dimension
We first discuss in detail the propagation of an acoustic wave, either longitudinal or transverse, in a direction normal to the interfaces of an infinite superlattice. The notation is shown in Fig. 7.1. Within the
Fig. 7.1 (a) Infinite superlattice. The layers are characterised by thickness dit density pt and elastic modulus (appropriate to longitudinal or transverse waves) Ct. The z axis is taken normal to the interfaces, (b) Acoustic-phonon dispersion curve for infinite superlattice when layers are impedance matched. Slopes of lines are u, where (a)
D
249
Layered structures and superlattices layers the displacement u satisfies the ID wave equation
We aim to solve these equations with the boundary conditions (see §2.1.1) that u and C du/dz are continuous at each interface. Since the system has translational periodicity D = dl+d2 we can introduce the wave vector component Q by means of Bloch's theorem: u(z + D) = Qxp(iQD)u(z) (7.2) The solutions of the wave equations in the two media at frequency co are a superposition of a forward- and a back ward-travelling wave. It is convenient to represent the solution within any one layer in two alternative forms, one with the phases referred to the left-hand end of the layer, and the other with the phases referred to the right-hand end. Thus we write u = a\ expCig^z - ID)~\ + b\ e x p C - i ^ z - ID)~] = a? e x p p ^ z - ID - dx)] + bf e x p E - i ^ z -IDdj\ lD^z^lD + dx
(7.3)
where q1=co/v1 The amplitudes are related by |«f> = Fi|«P> where we have denoted
(7.4) (7-5)
**>-(£)
(76)
A)
(7-7)
and the 2 x 2 matrix Fx is
SI
-
with / t = expOq^) In a layer of the second medium, the displacement is
(7.8)
u = d\ exp[ig2(z - ID - d j ] + e\ exp[-i^ 2 (z - ID - dj\ = df exp[i
Continuum acoustics: folded acoustic modes The boundary conditions at z = ID + d1 relate \uf} to |w^>: '1 1 where Z = Ciqi/C2q2 (7.12) is the ratio of the acoustic impedances of the media. Similarly the boundary conditions at z = (/ + \)D give
Note that the same matrices occur in (7.11) and (7.13) since the formalism presented so far is symmetric between +z and — z directions. The equations above combine to give a transfer matrix T defined by |ul+i> = 7 X > (7.14) where explicit evaluation gives Z-1)/152 -^i(Z-Z-1)f;1s2 \ 1 )/^ f^c2-±i(Z + Z-i)f^s2) K' ] with the shorthand notation c2 = cos(q2d2) (7.16) s2 = sm(q2d2) (7.17) It is readily verified from (7.15) that the transfer matrix is unimodular, detT=l (7.18) as is required for conservation of energy. We now apply Bloch's theorem. It follows from (7.2) that k + i > = exp(iQDK> (7.19) so that [T-exp(iGD)/]|ii{-> = 0 (7.20) where / is the unit 2 x 2 matrix. Equally from the equation relating \u^x)
to K>, [T-1-exp(-i6Z))/]^> = 0 (7.21) so that combining (7.20) and (7.21) [i(T + T~x) - I cos QD\\u\y = 0 (7.22) This holds for the amplitude \u[} in any cell /, and so the matrix must vanish. Hence we have / cos QD = i(T + T~x) = \l tr T (7.23) where the last step follows from (7.18) since T is a 2 x 2 matrix. Here 251
Layered structures and superlattices tr T denotes the trace of the matrix T. With (7.15) for T, this gives the explicit form of the dispersion equation: cos QD = cos qxdx cos q2d2 — [(1 + Z 2 )/2Z] sin qxdx sin q2d2
(7.24)
This equation is long established, having first been derived by Rytov (1956). Equation (7.24) has some elementary properties of interest. First, if Z = 1 the layers are impedance-matched, and the medium is effectively continuous acoustically. The equation then reduces to cos QD = cos(qldl + q2d2)
(7.25)
which has solutions co( -A + dA=±QD
+ nn
(7.26)
This is the equivalent of the 'empty-lattice' approximation of electron-band theory, since we have applied Bloch's theorem but without any reflections at the interfaces. As illustrated in Fig. 7.1(b), (7.26) corresponds to the folding of the acoustic-phonon dispersion curve into the reduced zone. Note, however, that the slope of the curve is an average of the velocities in the two media. For Z ^ l , w e have the inequality 1 + Z 2 > 2Z. If (7.24) is reorganised into a sum of two terms in cos(q1d1 ±q2d2) it is then readily seen that for some values of <x> the absolute value of the right-hand side is greater than 1. For these values of co there is no solution with real Q. Thus for Z ^ 1 there are some stop bands in the dispersion curve. It can also be seen that if qxdx and q2d2 are not commensurate, the dispersion curve is not, strictly speaking, periodic in co. That is, the magnitudes of the stop bands vary irregularly with co. These features are illustrated in Fig. 7.2. The above calculation, like the rest of this chapter, is concerned with superlattices in which the unit cell consists of two layers. The method can be extended to the case where the unit cell consists of N different layers. A range of theoretical results for this case has been presented by Djafari-Rouhani and Dobrzynski (1987). 7.1.2 Light scattering Inelastic light scattering has been used to investigate folded acoustic-phonon dispersion curves like those of Fig. 7.2. The first such observation was reported by Feldman et al. (1968), whose specimens were long-period SiC polytypes. We concentrate, however, on more recent results obtained on semiconductor superlattices. As always, inelastic light scattering essentially probes the dispersion relation along a vertical line 252
Continuum acoustics: folded acoustic modes
in the coq plane. By appropriate choice of specimen periodicity D and incident-light frequency the value of q can be chosen to be on a scale comparable with n/D, so that much of the superlattice Brillouin zone 0 < q < n/D is accessible to light scattering. Furthermore, substitution of typical values for D and the acoustic velocity vl shows that the excitation frequencies in Fig. 7.2 can correspond to tens of wavenumber units. This means that Raman scattering can be used, at least for the higher branches, rather than the more difficult technique of Brillouin scattering which is normally required for light scattering off acoustic modes. The first experimental results on the folded acoustic spectrum were obtained by Colvard et al. (1980). A good, more recent, example of a Raman scattering spectrum is shown in Fig. 7.3. The experiment was carried out in backscattering, and the polarisation selection rules are the same as those for backscattering off the surface of an ordinary specimen, as discussed in §2.5. Thus in (x, x) polarisation the LA phonons are seen, and there is no scattering by phonons in (x, y) polarisation. Jusserand et al. (1983, 1984a and b) carried out somewhat similar experiments. They used specimens with longer spatial periods of the order
Fig. 7.2 The folded-phonon dispersion curve of (7.24). The parameters are dl/d2 = 3/7, vjv2 = 0.856 and Z = 0.779. These are chosen arbitrarily to illustrate the stop bands and the nonperiodicity of the dispersion curve [after Babiker et al. 1985a] 30 -
20
10
B A
253
Layered structures and superlattices
of 20 nm, and they also varied the frequency of the incident light so as to probe a range of values of Q. By these means they were able to investigate much of the Brillouin zone. Positions of their scattering peaks, in comparison with the Rytov dispersion curve, are shown in Fig. 7.4. A general review of experimental results on folded acoustic phonons, and also on the folded optic phonons that form the subject of §7.2, is given by Klein (1986). The predominant experimental technique has been light scattering. 7.1.3 Green functions, oblique propagation and surface modes
Some developments of the Rytov-type theory presented above should be mentioned. By analogy with the detailed treatment of Brillouin scattering off isotropic specimens (§2.5), a full theory of lineshapes and strengths in light-scattering experiments should start with a derivation of the appropriate acoustic Green function. Babiker et al. (1985a) found the Green function for longitudinal or transverse phonons propagating normal to the interfaces in an infinitely extended superlattice. Their result was used in an approximate theory of the light-scattering cross section by Babiker and Tilley (1984) and Babiker et al. (1985b); the theory is approximate because of the use of the Green function for an infinite rather Fig. 73 Raman backscattering from a sample nominally consisting of a repeat unit of 4.2 nm GaAs and 0.8 nm Al 0 3 Ga 0 7 As. Inset shows dispersion curve corresponding to (7.24) for longitudinal acoustic phonons in Rytov's model, with crosses at experimental points. Arrows on main graph indicate peak frequencies corresponding to the superlattice period of 5.22 nm determined by X-ray diffraction [after Colvard et al. 1985]. 1000 •
800 • 600
a
80 120 Raman shift (cm 1 )
254
160
200
Continuum acoustics: folded acoustic modes
than a semi-infinite structure, because the optical propagation is treated in an 'average-medium' approximation rather than with the formalism to be presented in §7.3, and because possible ripple mechanisms for scattering were not included. The poles of the Green function are at the positions given by the Rytov model, (7.24), and therefore the scattering peaks are predicted to be at these positions. This is in agreement with experimental results such as those of Figs. 7.3 and 7.4. Apart from this more-or-less trivial result, the theory also predicts that line intensities should decrease with frequency, in general agreement with the experimental results. However, a detailed comparison of theory with experiment is not yet available. The theory of acoustic propagation at a general angle to the interface planes was given by Camley et al. (1983a) for s-transverse waves. The Fig. 7.4 Peak positions for Raman backscattering by folded LA phonons. Repeat unit: specimen (a), 5.6 nm GaAs/7.8 nm Al 0 7 7 Ga 0 23 As; specimen (b), 15 nm GaAs/10.7nm Al o . 36 Ga O64 As [after Jusserand et al. 1983].
i
5 -
0.2
0.4
0.6
0.8 6
1 1
Wavevector(10 cnr )
255
Layered structures and superlattices
derivation of the dispersion relation of the bulk modes is not much more complicated than was given above for normal incidence. Camley et al. go on to discuss surface and interface modes of a semi-infinite superlattice in contact with either vacuum or a bulk elastic medium. It is clear on general grounds that surface modes may occur at frequencies within the stop bands of dispersion curves like Fig. 7.2. It is found that they do indeed occur; for details the reader may refer to the original paper. Similar results are presented by Bulgakov (1985). It should be emphasised that in contrast to the Rayleigh and Stoneley waves discussed in §2.3 these surface modes occur in pure s-polarisation and are a consequence of the alternation of elastic properties depicted in Fig. 7.1. It may be noted that the authors also derive a Green function that in an appropriate limit should reduce to that given by Babiker et al. (1985a). The propagation of longitudinal plus p-transverse waves is discussed by Djafari-Rouhani et al. (1983). The elastic displacement in each layer is then described by four amplitudes rather than the two of (7.3) or (7.9), since there are two amplitudes for the longitudinal and two for the transverse component of the displacement. Consequently the transfer matrix is 4 x 4; the explicit form is given by Djafari-Rouhani et al. The authors present numerical illustrations for the Al/W superlattice. They show both the bulk modes of an infinite superlattice and the surface modes of a semi-infinite superlattice. Examples of their results are illustrated in Fig. 7.5. Although the theory has been given in detail for oblique propagation, experimental results are hardly available as yet. This is partly because of specimen shape: a semiconductor superlattice is essentially a thick composite film on a substrate. Because of the high refractive index, for the same reason as in the light-scattering experiments off ordinary semiconductors discussed in §2.5, experiments in which the incident and scattered light pass through the top face of the specimen are tantamount to normal incidence. It is possible for the incident light to be put in, or the scattered light taken out, parallel to the interfaces, but no results on acoustic modes in such a geometry have been reported. 7.2 Lattice dynamics: folded optic modes 7.2.1 One-dimensional model
The previous section was concerned with the folding of the acoustic-phonon dispersion curve into the mini-Brillouin zone resulting from the long period D. The specimens of most interest are based on III-V semiconductors such as GaAs. Since these materials are diatomic, they support optic phonons as well as acoustic phonons, and the optic 256
Lattice dynamics: folded optic modes
phonons in a superlattice are also folded. We begin this section with a detailed discussion of the simplest ID model, then discuss more realistic models and experimental results. The experiments, which are reviewed in more detail by Klein (1986), are again mainly Raman scattering. The model to be discussed is summarised in Fig. 7.6. It is assumed that the interatomic spacing a and elastic constant C have the same values in the two components of the superlattice, the difference between the two components being the change from mass mx to m2. The model may be seen as a simple description of a GaAs/A^Gax _xAs superlattice, m0 being the As mass. The dynamical behaviour of an infinite ID diatomic lattice was summarised in §2.1.2. The dispersion equation is given in (2.17), with further characterisation in Fig. 2.2 and Table 2.1. For the superlattice we take solutions in cell / of the form of (7.3) and (7.9). Now, of course, qx and q2 are related to co by (2.17) with appropriate mass values substituted. There is no need to restrict attention to frequency intervals in which qx and q2 are both real; it will be seen in fact that there are phonon modes
Fig. 7.5 Bulk continua and surface bands for longitudinal plus p-transverse modes in the Al/W superlattice (full line, Al on surface; broken line, W on surface): (a) d\\ = d^; (b) d&\ = 5d w . The surface-layer thickness is assumed to be the same as that of a corresponding bulk layer. The variables are dimensionless with cf(Al) denoting the transverse velocity of sound in Al, and D = d^\ + d\y [after Djafari-Rouhani et al. 1983].
Q
Layered structures and superlattices of the superlattice in which only one of ql and q2 is real. The algebra is the same whether or not q1 and q2 are real. The amplitudes at the left and right ends of a layer are still related by the matrices F1 and F2 of (7.7). In §7.1, the amplitudes in the two constituents of cell / were related by the two boundary conditions at the interface to yield (7.11) and (7.13). As was seen in §2.1, the latticedynamical equivalent is provided by the equations of motion of the atoms on either side of the interface. Thus the equations of motion of the masses indicated by arrows in Fig. 7.6 provide the analogues of (7.11) and (7.13). The results are
1
' W>-( '
(7.27) OC2S2
and 1
1 a2s2"1
1
a11s l
1 |wf> a 2 s 2i
(7.28)
where st = exp(i^a)
i = 1,2
(7.29)
and OL1 and a 2 are the ratios of the displacements of the two masses in the relevant component, given by (2.16): 2C — m0co2
a)
i=l,2
2C - m^2
(7.30)
Equations (7.7), (7.27) and (7.28) combine to give a transfer matrix T as defined in (7.14). The detailed algebra gives
Fig. 7.6 Microscopic model for a diatomic superlattice in the region occupied by unit cell /. The masses are coupled by nearest-neighbour forces with spring constant C.
2
X . — /-I
258
O
•
- 2nx a •
. 2n2a
Component 1
Component 2
O •
O •
O:x cell /
O x O
x
Lattice dynamics: folded optic modes
It can be confirmed from these expressions that det T = 1. The dispersion relation continues to be given by (7.23), which can be brought into the explicit form
x sinO?^) sin(g2d2) (7.33) This result was first given by Jusserand et al. (1984a), and the explicit derivation by a Green function technique is given by Djafari-Rouhani et al. (1985). Equation (7.33) contains a considerable amount of information, which we now explore. In the continuum limit qxa« 1 and q2a « 1 it reduces to (7.24), as it should. The result of a numerical solution of (7.33) for a specific choice of parameters is shown in Fig. l.l{b). In order to interpret this result, we include in Fig. 7J(a) the dispersion curves for the two bulk media. This enables us to divide the frequency axis into six regions, as indicated on the figure. As a development of Table 2.1, Table 7.1 shows the character of the bulk wavenumbers qx and q2 in each of these regions. Comparison with Fig. 1.1 (b) shows that where qx and q are both real, in regions I and IV, the superlattice dispersion curve shows broad pass bands and narrow stop bands. This is similar to what was found for the continuum description in §7.1. In regions III and V, however, where only one of qx and q2 is real, the pass bands are narrow and the stop bands broad. A qualitative reason for this can easily be seen from (7.33). In region V, for example, q1 = \y. The trigonometric functions then become hyperbolic: c o s ^ ^ ) = cosh();1d1), sm(qldl) = \smh(yld1), tan(^xa) = i tanh(};1a). Thus the right-hand side of (7.33) is essentially of the form Ax cos(q2d2) + A2 sm(q2d2), with |X 1 |»1 and |y4 2 |»l. This expression can be rewritten as B cos(q2d2 + 0), with B » 1. Thus as co moves up through region V (7.34) has solutions for Q only in the very narrow frequency bands where \cos(q2d2 + 0)|« !/#• A n analogous situation occurs for region III. The phonon modes in regions III and V have an oscillatory spatial dependence in the medium with real q and an exponentially decaying spatial dependence in the other medium. They are usually called confined phonon modes. The frequency varies only slightly with Q, and is therefore the same as would be found in a single slab of the real-g medium embedded in the other medium. The above model is the simplest possible for folded optic modes but we shall see that it is in, at least, qualitative agreement with Ramanscattering results. A much more complete theory using 3D lattice dynamics with realistic values for the semiconductor parameters was given by Yip 259
Layered structures and superlattices Table 7.1. Frequency intervals of Fig. 7.7 Frequency interval
Superlattice dispersion curve
I
R
R
II III
C R
C C
IV
R
R
V
I
R
VI
I
I
Broad pass bands, narrow stop bands No solutions Narrow pass bands, broad stop bands Broad pass bands, narrow stop bands Narrow pass bands, broad stop bands No solutions
R = real, C = (n/2a) + iy, I = pure imaginary.
and Chang (1984). Their results have the same general form as shown in Fig. 7.7. The most important physical feature that is omitted from the simple model calculation is the influence of the long-wavelength Coulomb Fig. 7.7 (a) Bulk dispersion curves, (2.17), for two constituents of a superlattice with m 1 /m 0 = 0.93, m 2 /m 0 = 0.70. Frequency axis is co/a)u where col = [2C(mQl +m 1 ~ 1 )] 1 / 2 is the zone-centre optic-phonon frequency in medium 1. (b) Dispersion curve for a superlattice of these constituents with nl=n2= 10, i.e. 10 unit cells in each layer [after Albuquerque et al. 1988].
0.2 0.4 0.6 0.8 1.0 1.2 1.4 n/2 qa {a)
260
0
0.5
1.0
1.5 QD (b)
2.0
3.0
Lattice dynamics: folded optic modes interaction on the optic modes. In a bulk polar medium, this interaction is responsible for the LO-TO splitting, and is crucial to the theory of the dielectric response. Although Yip and Chang include the Coulomb interaction by a perturbation method, there is as yet no full and clear account of the matter. Most experimental groups have so far made comparison with the simple model. 7.2.2 Light scattering As with the folded acoustic modes, the main experimental technique that has been applied to study the folded-optic mode spectrum is Raman scattering. The calculation presented above is concerned with phonons propagating normal to the interfaces, so the relevant experimental results are from 180° backscattering at normal incidence. The selection rule for this geometry is that only longitudinal modes contribute to the cross section, so the data involve folded LO phonons. Figure 7.8 shows Raman spectra obtained by Jusserand et al. (1984b, 1985) from a number of GaAs/Al0 3 Ga 0 7As superlattice specimens. The peaks labelled 1 to 4 Fig. 7.8 Raman spectra in the LO phonon region from four samples consisting of repeat units having n monolayers of GaAs and ri monolayers of Al 0 3 Ga 0 7 As. The (n, ri) values from top to bottom are: (6,4), (9,9), (12,7), (17,12) [after Jusserand et al. 1984b]. T = 8 0 K X,
514.5 nm
z(x,y)z
S2
300
295 290 285 280 Frequency shift (cm"1)
275
261
Layered structures and superlattices
are the highest four confined states; peak A is assigned to an AlGaAs LO mode. The peaks correspond to LO modes confined in GaAs by AlGaAs 'phonon barriers', so that their positions should depend on the GaAs monolayer number n but not significantly on the AlGaAs monolayer number. Figure 7.8 shows the variation with n of the peak positions compared with theoretical curves derived from (7.33). The theory depends on the value of the AlGaAs LO phonon frequency; two possibilities are shown on Fig. 7.9. Figure 7.9 shows that the line intensity decreases with increasing folding number, as it does for folded acoustic phonons, Fig. 7.3. A theoretical account involves evaluation of the intensities from the relevant Green function theory by S. R. P. Smith (unpublished), and the resulting intensities do indeed decrease with folding number in agreement with the experimental results. Raman-scattering results relevant to optic modes that are not propagating normal to the interfaces were published by Zucker et a\. (1984, 1985). They arranged that the superlattice specimen was an optical waveguide by surrounding it with AlGaAs cladding layers. By this means they were able to guide out the scattered light in the plane of the layers, so that they had a 90° scattering geometry. The positions and polarisation Fig. 7.9 LO phonon frequencies versus n for (GaAs)n(Al0 3 Ga 0 7As) superlattices. Circles give experimental results. Lines give results of calculations based on (7.33). Full line for AlGaAs LO frequency corresponding to 280 cm" 1 , broken line to 286 cm" 1 [after Jusserand et al. 1984a].
S 285 -
10
262
20
Optical properties dependence of their peaks were consistent with a simple model in which the specimen is treated as a set of isolated GaAs layers. Each layer will then sustain Kliewer-Fuchs modes as described by (6.62) and (6.63); Zucker et al. were able to fit their results by taking the expressions for a GaAs layer in vacuum. The theoretical significance of this simple result is not altogether clear. 7.2.3 Surface modes Equation (7.33) and the subsequent discussion were concerned with the bulk modes of an infinite superlattice. As in continuum acoustics, surface modes may occur in stop bands, such as those of Fig. 1 .l(b). A brief discussion of these surface modes for the ID model is given by Djafari-Rouhani et al. (1985). Their calculation is analogous to that of Wallis (1957) for the surface mode of a semi-infinite ID diatomic crystal, which was outlined in §2.1. 7.3 Optical properties 7.3.1 General results The complete formal theory of optical propagation in an infinite periodic layered medium in which the constituents are isotropic or cubic was given by Yeh et al. (1977) and Yariv and Yeh (1977). An account is also given by Yariv and Yeh (1984). The theory is very similar to the continuum acoustics of §7.1. There is, however, the simplification that in optics the separation between transverse and longitudinal modes is maintained in oblique incidence. Thus the theory for waves travelling in an arbitrary direction through the superlattice is not much harder than that for waves travelling normal to the interface, and we proceed straight away to the general case. As indicated in Fig. 7.10, the notation is similar to that used for the acoustics calculation. Each medium is characterised by an isotropic dielectric function e x, s2 or equivalently refractive index rjx = e} /2 , rj2 = s\12. The z axis is taken normal to the layers, as before, and the x axis is chosen so that the propagation vector lies in the xz plane. As usual, therefore, we can distinguish s-polarisation, with E in the y direction, from p-polarisation, with E in the xz plane. We start with s-polarisation, for which the formalism of §7.1 applies with minor modifications. All fields are taken to have x- and ^-dependences proportional to exp(ig xx — ia>t). The E field in a layer of medium 1 can then be written in either of the forms of (7.3), where now ql is given by q\ + ql = sl(a2lc2
(7.34) 263
Layered structures and superlattices Equation (7.5) relates the amplitudes at the two ends of the layer, with the phase matrix Fx given by (7.7) and (7.8). The E field in medium 2 is written as (7.9), with (7.10) relating the two forms. The boundary conditions at an interface are that Ey and Hx are continuous, that is, Ey and dEJdz are continuous. Thus (7.11) and (7.13) are replaced by
wf> = <
t>
(7.35)
and (7.36) where ' = 1,2
(7.37)
Equations (7.5), (7.10), (7.35) and (7.36) combine to give a transfer matrix T: T = Q;1Q2F2Q;iQ1Fl
(7.38)
Explicit evaluation of T requires patience rather than insight; for example, the diagonal elements are Tn = / i [ / 2 f a i +q2)2 - / 2 ' f o i -
(7.39)
and T22 =f;1U2-1(qi+q2)2 -fiiqi-qifV^qi (7.40) The application of Bloch's theorem as given in §7.1 now shows that the dispersion relation is cos(QD) = cos(q1d1)cos(q2d2) - gs s i n ^ ^ J sin(q2d2) Fig. 7.10 Notation for optical calculation.
264
(7.41)
Optical properties where ( 7 - 42 ) The calculation for p-polarisation is similar to that just given. As noted in §6.3, the boundary conditions for s-polarisation can be converted into those for p-polarisation by the replacement q2/qi ->£2<7i/ei42- Thus the dispersion relation is given by (7.41), with gs replaced by GP = i(e 2 0i/£i02 + £i42/£24i)
(7.43)
Equation (7.41) gives the dispersion relation co = co(qx, Q); qx and co enter via (7.34) for qx and q2. A very full discussion of the implications when sx and e2 are positive and independent of frequency is given by Yeh et al. (1977) and Yariv and Yeh (1977). As for the acoustic case, band gaps occur at the points QD = nn. It is convenient now to work in an extended-zone representation; a typical curve of co versus Q for either polarisation is sketched in Fig. 7.11 (a). An alternative representation consists of drawing contours of constant frequency in the qxQ plane, as shown in Fig. 7.1 l(b). The contours for s- and p-polarisation are different, except that they touch in the direction qx = 0. This corresponds to propagation normal to the interfaces, for which s- and p-polarisation are indistinguishable. The band gaps at Q = n/D have the effect that the contours in thefirstzone are pulled out towards the zone edge; by analogy with electron-band theory this is a consequence of the gaps shown in Fig. 7.11 (a). The contours may be compared with those drawn for a conventional uniaxial medium, as shown in most optics texts (see, e.g., Lipson and Lipson 1969). In that case the contours corresponding to the ordinary wave are circles, while those corresponding to the extraordinary wave are ellipses. For the superlattice neither set of contours is circular, so in a sense both polarisations correspond to extraordinary propagation. Furthermore, for a frequency and direction corresponding to a gap region, there can be none or only one propagating mode rather than the two that are always found in a conventional medium. Although the theoretical treatment for positive and constant ex and s2 was given in complete and definitive form by Yariv and coworkers, there has been little experimental investigation. The case when s 1 and e2 exhibit dispersion is also clearly of considerable interest, not least for semiconductor superlattices. The theoretical development, down to (7.41), continues to apply, but a full discussion of the implications is lacking as yet. Similarly, there is room for much more experimental work. A useful simplification of (7.41) for the very long-wavelength limit was given by Raj and Tilley (1985). At low frequencies, as for a semiconductor in the reststrahl region for example, the wavelength is much longer than 265
Layered structures and superlattices
the periodicity D. Thus the inequalties qxD«l, QD«1, qxD«\ and q2D « 1 may be expected to hold; in terms of Fig. l.l\(b) we are discussing contours well inside the first Brillouin zone. In that case all the trigonometric functions in (7.41) can be expanded. Retention of the first two Fig. 7.11 {a) Dispersion curve OJ versus Q for fixed qx for optical propagation in either polarisation. In an extended-zone scheme, band gaps appear at Q = nn/D. (b) Constant-frequency contours in the qxQ plane. {a) IDD/C
n/2
it)
266
Optical properties terms for the cosines and the first terms for the sines leads to the simplified equations ZxxiQ1 + <7x) = u>2/c2 ZxxQ
2
e
1(
2 2
+ zz ll = u> /c
s-polarisation
(7.44)
p-polarisation
(7.45)
where exx = (eidi+s2d2)/D e
l=
e 1(
(7.46)
E1
z~z ( r ^i + i d2)ID (7.47) The same results, with alternative derivations and different physical discussion, were obtained at about the same time by Agranovich and Kravtsov (1985) and by Liu et al. (1985). The argument presented by the former authors is particularly illuminating. At long wavelengths the boundary conditions that Ex and Dz are continuous means that these field components have the same values over many periods of the superlattice. Thus we write Exi= E
_J
i=l,2
(7.48)
where Ex, Dz are the constant values of the fields. The other field components are given by
I= 1 2
'
E"-1% ^zi
8
i
—
U
(749)
zi
The spatial averages Dx = (d1Dxl + d2Dx2)/D and Ez are then given by Dx = sxxEx _ (7.50) x * Ez
=
£
zz ^z
with the dielectric-tensor components given by (7.46) and (7.47). Equations (7.44) and (7.45) show that in the long-wavelength limit the superlattice behaves as a uniaxial medium with the averaged (or effective) dielectric function obtained from (7.46) and (7.47). Raj and Tilley (1985) discuss the reststrahl region of GaAs/Al^Ga! _ xAs as a particular example. The frequency dependence of sxx and szz is quite complicated since Al^Ga^^As exhibits two-mode behaviour (see Problem 6.1), while GaAs has a single TO resonance. For this system, (7.46) has been used by Maslin et al. (1986) to give an account of normal-incidence far-infrared reflectivity measurements. The experiments were carried out by dispersive Fourier transform spectroscopy (DFTS), which as explained by Birch and Parker (1979) is a technique for making simultaneous measurements of the amplitude and phase of the complex reflectivity. Experimental results and theory are compared in Fig. 7.12. The theoretical curves represent a leastsquares fit to the data. The theoretical expressions are given by the usual 267
Layered structures and superlattices
classical optics of a multilayer system, with the superlattice described by (7.46). The dielectric functions ex and e2 appearing there are taken in the forms of (5.60) and Problem 6.1, and the parameters in those expressions are found from the fit to experiment. Equations (7.46) and (7.47) can be used together with standard results for optical propagation in anisotropic media to give an account of bulkpolariton propagation in the long-wavelength limit. This is done by Raj et al. (1987). 7.3.2 Green functions
Linear-response theory has been used by Babiker et al. (1987a) to evaluate the electric-field Green functions for an infinitely-extended two-component dielectric superlattice, as depicted in Fig. 7.10, with retardation effects included. Briefly, the calculation involves using Maxwell's equations to write down expressions for the electricfieldvectors in each component layer of the superlattice in the presence of an externally applied polarisation field Pext. The standard electromagnetic boundary conditions are applied at each interface, just as in the corresponding Fig. 7.12 Far-infrared reflectance data for a specimen including a superlattice comprising 60 periods, each composed of 5.5 nm of GaAs and 17 nm of Al 0 3 5 Ga 0 65 As. The insert shows the specimen structure with thicknesses given in jim. Regions 1 and 3, Al 0 3 5 Ga 0 65 As; region 2, superlattice; region 4, GaAs substrate. In the main part of the figure, crosses are experimental points and the curves are theoretical [after Maslin et al. 1986]. 1.0
0.0
268
250
300 350 Wavenumber (cm 1 )
400
450
Optical properties
homogeneous calculation in §7.3.1, and the periodicity property of the superlattice is utilised through Bloch's theorem as in §7.1. The required Green functions, which are of the form <<£^(g); £v(Q'))>a>» are then obtained from the linear response of the electric-field components to the polarisation Pext, in accordance with general results in the Appendix. Here \i and v denote Cartesian components, and Q and Q' are components of wave vector in the z direction. For an infinitely extended superlattice the Green functions are in fact all proportional to S(Q — Q'). It is found that the Green functions with \i and v both equal to y describe s-polarised modes; they have poles corresponding to the dispersion relation (7.41). Likewise those Green functions with \i and v equal to either x or z describe p-polarised modes. The general Green function results are rather complicated and will not be given here. However, as shown by Babiker et al. (1986c, 1987a), they simplify considerably in the low-frequency limit where retardation effects become unimportant. Some specific applications to light scattering from plasma modes are quoted in §7.4. 7.3.3 Surface modes
It has been brought out in previous sections that surface modes can appear on a semi-infinite superlattice in the stop bands of the folded dispersion curve. A striking illustration occurs with s-transverse acoustic modes, where as mentioned in §7.1 surface modes are predicted on superlattices although they do not occur on bulk specimens. Similarly, surface optical modes can occur simply as a result of the periodic alternation of the refractive index, for example in s-polarisation and with real refractive indices. Surface modes of this kind were already discussed by Yeh et al. (1977). However, in optics surface modes can also occur because one or both component media has a negative dielectric constant, and thus is surface-active. The full range of possibilities arising from these two different mechanisms has not been fully discussed in any specific case. Surface polaritons of the second kind in the reststrahl frequency region have been discussed within the long-wavelength approximation of §7.3.1 by Raj and Tilley (1985) and Raj et al. (1987). Within that approximation the superlattice is described as a uniaxial medium with dielectric tensor components given by (7.46) and (7.47). The surface-polariton dispersion equation is therefore given by (6.42), with the simplification that the x and z direction are principal directions of the dielectric tensor. The calculated dispersion curve for a semi-infinite GaAs/Al^-Ga^^As superlattice bounded by vacuum is shown in Fig. 7.13. Surface polaritons of this kind must satisfy the localisation condition exx<0. In Fig. 7.13 a distinction is made, following Hartstein et al. (1973), between modes with 269
Layered structures and superlattices
szz < 0, called real-excitation surface polaritons, and modes with ezz > 0, called virtual-excitation surface polaritons. In general, modes of the former kind continue to exist as qx^> oo, while the latter terminate at a finite value of qx. In addition to the dispersion curve, Raj et al. present theoretical results for the ATR spectrum of a finite superlattice. The spectrum shows considerable structure, including guided-wave as well as surface polaritons. A range of general results for the p-polarised polaritons of a finite superlattice were given by Haupt and Wendler (1987). In addition to the dispersion curves, they present a perturbation calculation in which the damping of the surface modes is found to first order in the damping constants F appearing in the expressions for the dielectric functions of the constituent media. Their most important conclusion is that just as for the homogeneous film discussed in §6.3, the surface mode of higher frequency has a long propagation distance. Thus the superlattice supports a long-range surface polariton. 7.3.4 Limits of validity of theory The results and discussions of the previous parts of this section have been based on the assumption that each component of the superlattice Fig. 7.13 Surface-polariton dispersion curves a> versus qx, within the long-wavelength approximation, for the interface between a GaAs/Al 0 1 4 Ga 0 86 As superlattice and vacuum. The superlattice is described by (7.46) and (7.47) with dl=d2 and values for the reststrahl parameter taken from Maslin et al. (1986). Full line, realexcitation surface polaritons; broken line, virtual-excitation surface polaritons; dotted line, vacuum light line [after Raj et al. 1987]. 450 -i
400-
350u)/2nc (cm"1)
300-
250200-
250
300
350 qx/2n (cm-1)
270
400
450
500
Single-electron states is characterised by a dielectric constant £, and it is assumed either implicitly or explicitly that e has the same value as it would have in the corresponding bulk medium. However, as mentioned in §5.4, the theory of the bulk dielectric function, for example in the reststrahl region, is essentially the theory of the response of the macroscopic polarisation P to the macroscopic electric field Emac. In the course of the calculation, it is necessary to express the local electric field Eloc in terms of P and Emac. For superlattices with thick layers, as used for Fig. 7.12 for example, it is probably justified to use the bulk e since the expression for Eloc is little altered from the bulk. When the component layers are only a few monolayers thick, however, the expression for Eloc must be modified and the assumption of a bulk e is not justified. The calculation that is then needed is equivalent to the inclusion of the Coulomb interaction in the lattice dynamics of §7.2, and as mentioned there no clear account has yet been given. It is for this reason that although Maslin et al. (1986) present experimental data for a specimen with a spatial period of four monolayers, they do not give any comparison with theory. 7.4 Single-electron states 7.4.1 Envelope-function approximation As was made clear at the beginning of this chapter, the main reason for the very large investment in MBE technology has been to gain more control over electronic properties than is possible with bulk specimens. Consequently electron states have been much more extensively studied than phonons or optical properties. It is beyond our scope to go into much detail; this section is therefore restricted to an introduction to the main ideas, presented so as to bring out the similarities to acoustic and optic properties. The reader who wishes to pursue the topic further should look elsewhere, and could start with the chapters by Kroemer and Bastard in Chang and Ploog (1985). The simplest picture of a semiconductor is of a conduction band and a valence band separated by a gap £ g. Correspondingly, a useful picture for electrons in a superlattice is shown in Fig. 7.14. 'Wells' of width dx of material 1, which has gap El9 are separated by 'barriers' of width d2 of material 2, which has gap E2. This picture shows the bottom of the conduction band and the top of the valence band, with degeneracy neglected. It is appropriate to GaAs/A^Ga^^s with x<0.3, among other materials, since GaAs and AlxGax_xAs are both direct-gap semiconductors, with both valence-band maximum and conduction-band minimum at the F-point (the centre of the Brillouin zone). AlAs and Al/ja^^As with x >0.3, on the other hand, are indirect-gap materials, 271
Layered structures and superlattices with conduction-band minimum at the X-point. Electron transfer across the interfaces is a more complicated process in such materials. Of the quantities defined in Fig. 7.14, Ex and E2 are known, and dl and d2 are known from the growth conditions. However, the band offset, Ec say, is not known a priori and has to be determined for each superlattice system. In some systems, Ec is negative. We now concentrate on motion of a single electron in the potential described by the upper half of Fig. 7.14. Bastard (1981) showed that provided the individual components are more than a few monolayers thick the motion can be described by means of the envelope function F(r) that describes the modulation of the electron wavefunction from one atomic site to the next. To be specific, the wavefunction is written
(7.51)
where wo(r) is the Bloch wavefunction corresponding to the bottom of the conduction band, satisfying wo(r + a) = wo(r), where a is a unit vector of the lattice. Provided F(r) varies only slowly on the scale of a, it satisfies Schrodinger's equation for a particle with the effective mass m appropriate to the conduction band. Bastard (1981) further shows that in the approximation that the bands are parabolic, E = h2q2/2m, the boundary conditions at the interface between two component layers are F1=F2
(7.52)
x
l
m; dFJdz = m2 dFJdz
(7.53)
Equation (7.53) implies continuity across the interface of the probability current density J = (h/2im)(F*dF/dz - FdF*/dz)
(7.54)
Further discussion of the envelope-function approximation is given in the chapter by Bastard in Chang and Ploog (1985). 7.4.2 Transfer matrix and dispersion relation The description of the electron motion now reduces to solving the Kronig-Penney model with the boundary conditions (7.52) and (7.53). Fig. 7.14. Simple model for electronic properties of a superlattice.
A V
272
Single-electron states For convenience we simplify and redefine the notation to that shown in Fig. 7.15. The calculation is very similar to those given in previous sections. For electron energy E > Vo, the wavefunction is a sum of forward- and backward-running waves in each medium: F(z) = ax cxpliq^z — ID)] + bx exp[ — \qx(z — ID)] lD
+ d^
(7.55)
F(z) = c, exp[i<72(z - ID - dj] + dx exp[-i^f 2 (z - ID - dj] /D + d 1 < z < ( / + l ) Z ) (7.56) where E = h2q2l2m1
(7.57)
2 2
E-V0 = h q /2m2 (7.58) As in §7.1.1, the boundary conditions are used to find a transfer matrix T relating (al + ubl+1)to (ah bx). Application of Bloch's theorem then gives the dispersion relation in the form of (7.23), and substitution of the explicit form of T (Problem 7.4) gives the result cos(gD) = c o s ^ d j cos(q2d2) — \(C 4- C~x) s i n ^ ^ J sin(^f2rf2)
(7.59)
where C = m2qjm1q2
(7.60)
For electron energy E < Vo, the wavefunction in the barrier regions is a sum of real exponentials. Formally, q2 is replaced by i/c2, where Vo - E = h2K\l2m2
(7.61)
The dispersion relation becomes cos(gD) = c o s ^ ^ i ) cosh(K:2rf2) — \(B — B~l) s i n ^ ^ J sinh(K:2^2)
(7.62)
where B = m2ql/m1K2
(7.63)
The discussion of (7.59) and (7.62) proceeds along familiar lines, and the analogy with the lattice dynamics of §7.2.1 is particularly close. For E > Vo both qx and q2 are real; this is a region of wide pass bands and narrow stop bands. For E < Vo ql is real and q2 is imaginary. We now have bound states in the well regions 1 coupled by evanescent waves in Fig. 7.15. Notation for Kronig-Penney calculation. v — v r
—
r
1
2
n
v=o-
ID
273
Layered structures and superlattices
the barrier regions 2. For K2d2 large, the coupling is weak: this is the tight-binding limit, with almost-flat bands at the bound-state energies of the wells. These comments are illustrated by Fig. 7.16. For the well depth chosen, the isolated well has three bound states. The barrier width is such that the well states are only weakly coupled, so for E < Vo we have very flat bands. For E > Fo, on the other hand, the stop bands are very narrow. Figure 7.16 may be compared with Fig. 7.7, which shows similar effects for lattice dynamics. It was mentioned at the beginning of this chapter that the term 'superlattice' is often used with a more restricted meaning as far as electronic states are concerned. In this context 'superlattice' denotes a specimen in which the barriers are relatively thin so that there is significant overlap between the wavefunctions of bound states in adjacent wells, and the E-Q curves arising from bound states therefore show significant dispersion. Conversely, the term 'multiple quantum well' (MQW) usually denotes a specimen in which wavefunction overlap is negligible and the E-Q curve is flat. Thus Fig. 7.16 corresponds to the MQW case. Preliminary experimental results related to the band structure for electronic motion perpendicular to the layers have been reported by Duffield et al. (1986). As indicated in Fig. 7.17, they carried out cyclotron resonance experiments on a superlattice specimen with the magnetic field Fig. 7.16. Electron bands in the envelope-function approximation for d1/d2 = 0.2 and im^lVJh2 = 80. (a)m2/m1 = 1, (b)m2/m1 = 2.0, (c)m2/m1 = 0.5 [afterR. N. Philp unpublished]. 120r 100
80
r
V 60 40
20
0
274
0.5
1.0
1.5 QD
2.0
2.5
3.0
Single-electron states applied in the plane of the layers. As shown there, the electron orbit traverses the layers, and the cyclotron-resonance frequency is coc = eB/mc, where mc is given by the standard expression for an anisotropic material (see Kittel 1986): mc=(h2/2n)dS/dE
(7.64)
and the derivative is the rate of change with electron energy E of the orbit
to
120
•
—
.
—i
•
100
H
l
'
1
80
60
I
40
20 n
-
0
•
.
i
i
i
i
0.5
1.0
1.5 QD
2.0
2.5
3.0
I
275
Layered structures and superlattices
area S in g-space. Loosely speaking mc is the geometric mean of the effective masses parallel and perpendicular to the layers. In subsidiary experiments in which B is perpendicular to the layers, the effective mass for motion in the plane is measured. Combination of the two measurements therefore yields the effective mass for motion perpendicular to the interfaces, which is determined by the curvature of the appropriate E-q curve in dispersion graphs like Fig. 7.16. The experimental method consisted of measurement of far-infrared transmission through the specimen by means of DFTS spectroscopy, which was mentioned in §7.3.2. Typical results are shown in Fig. 7.18. As explained, the combination of resonance frequencies for the two orientations of magnetic field yields a value for the effective mass perpendicular to the interfaces. Duffield et al. determine this mass for five different values of concentration x, obtaining results that are in satisfactory agreement with the envelope-function approximation. 7.5 Plasmons
This section is concerned with the coupled plasma oscillations of the electrons (or holes) in superlattices and MQWs. A simple theoretical Fig. 7.17 Magnetic-field orientation and electron orbit in the cyclotron-resonance experiment of Duffield et al. (1986).
276
Plasmons
description, which is applicable if the well thickness (d1 in the notation of Fig. 7.15) is sufficiently small, is to treat the charge layers as being 2D sheets separated by a dielectric medium of thickness d2. This approach is discussed in §7.5.1. Then in §7.5.2 we present the theory based on taking the charge layers to be of finite, rather than infinitesimal, thickness. In both cases, the calculations may be carried out either microscopically (e.g. by starting from a Hamiltonian and using RPA) or macroscopically (e.g. by hydrodynamic theory or by a description based on solving Maxwell's equation); this is a similar situation to the case of plasmons in a single medium as discussed in §5.3. For simplicity the account given here will emphasise the macroscopic approach. Light-scattering studies have provided verification of many of the theoretical results for the plasmon modes of a superlattice (or MQW), and we give an account of this work in §7.5.3. 7.5.1 Two-dimensional charge-sheets model
As discussed already in §5.3, the bulk 3D plasmon frequency at long wavelengths has a constant value cop, while the plasmon frequency Fig. 7.18 Inverse of transmission coefficient in cyclotron resonance experiments on an n-type specimen at 70-80 K. The curves peak (minimum transmission) at the resonance frequency. Applied field values are marked [after Duffield et al. 1986].
1.25-
1.00
1.25 -
1.00 50 100 Frequency (cm"1)
277
Layered structures and superlattices
of a single 2D layer is proportional to the square root of the wavevector 4l|. In the case of an infinitely extended periodic array of equally spaced 2D electron layers, the plasmon mode wasfirstcalculated by Fetter (1974) and subsequently studied by Das Sarma and Quinn (1982) and Bloss and Brody (1982), indicating a behaviour that lies between the above two extremes. To rederive their results in a form appropriate to long wavelengths we employ the model and notation of Fig. 7.19. Maxwell's equations may be utilised within each dielectric medium (A) subject to the standard electromagnetic boundary conditions at the charge-layer interfaces. The transverse electric field within the dielectric medium of cell n may be written as £?>(*, z) = exp[i(^x - (ot)]{a(n) exp[iqz(z - nD)] + bin) exp[-i^ z (z - nD)-]} (7.65) where ain) and bin) are the amplitudes for the forward- and backwardtravelling waves, and the condition for the electric field to satisfy the wave equation gives qz = (eco2/c2-q2x)^2
(7.66)
Without loss of generality we have taken the in-plane wavevector qy to be in the x direction. Using V x H = £oe(dE/dt), the magnetic field in medium A of cell n is obtained as
H{;\x, z) = (eoea>/qz) exppfayc - cot)]{a(n) exp[iqz(z - nD)] -b^Qxpl-iqz(z-nD)]} (7.67) The usual boundary conditions, that the electric field Ex is continuous across any interface and the discontinuity in the magneticfieldHy is equal to the current density at any interface, can now be applied at the interface z = nD. The complete solution is obtained noting that Bloch's theorem Fig. 7.19 Schematic illustration of a superlattice of equally spaced 2D charge layers separated by a dielectric medium (A). z= (n-l)D
n-\
278
Unit cell
D
nD
«+l
n+2
Plasmons implies (e.g., see (7.2)) a(n+1) = Qxp(iQD)ain)
b(n + 1) = exp(iQD)b{n)
(7.68)
where Q is the component of the excitation wavevector in the z direction. This leads to two linear homogeneous equations involving a(n) and b{n\ and the solvability condition is obtained as (see Problem 7.5) 2co2[cos(QD) - cos(qzD)] - Q2qzD sin(qzD) = 0
(7.69)
Here Q denotes a characteristic frequency parameter defined by Q = (noe2/sosm*D)1/2
(7.70)
where n0 is the areal density of electrons (of effective mass m*) in each 2D sheet. Equation (7.69) could equally well have been obtained by transfer matrix methods. The dispersion relation (7.69) simplifies in the limit q2. »sa>2/c2, which corresponds to neglecting retardation effects, because we then have qz = iqx. Substitution into (7.69) gives
_ j
qxDsmh(qxD) \^2 [2[cosh(^D)cos(eD)]J
l
'
}
in agreement with calculations by Fetter (1974), Das Sarma and Quinn (1982) and Bloss and Brody (1982) using different methods. It is easily verified from (7.71) that, except when cos(gD) = 1, we have co proportional to qx at long wavelengths (such that qxD«l). The behaviour of the plasmon frequency as a function of qx in the nonretarded regime is illustrated in Fig. 7.20 for three different values of QD. There are also additional solutions of (7.69) at higher frequencies such that eco2/c2 >; ql, when retardation effects become important. These have the characteristic of the perturbed photon line folded back in a reduced zone scheme (e.g., see Constantinou and Cottam 1986). Next we consider the plasmon modes of a semi-infinite superlattice, having the same charge layer structure as in Fig. 7.19 but occupying just the half-space z ^ 0. The region z > 0 is assumed to be filled by a medium of dielectric constant es, and we define the ratio rs = sjs. The surface z = 0 between the outermost cell of the superlattice and the external medium is different from all the other interfaces within the superlattice, and so the electron density nOs associated with its charge layer may be different from n0 (or even zero). On defining /as = nOs/no, we have two surface parameters rs and /*s as in Constantinou and Cottam (1986). In earlier calculations for semi-infinite arrays of 2D charge layers (Giuliani and Quinn 1983; Giuliani et al. 1984; Jain and Allen 1985a) a more restricted boundary condition was employed in which the only parameter was rs, whilst fis was arbitrarily taken equal to one. 279
Layered structures and superlattices The calculation of bulk and surface plasmon modes can be made by a direct generalisation of the infinite superlattice case. The electric and magnetic fields in each region of space, including z > 0, can be written down as in (7.65)-(7.67). The usual boundary conditions can then be applied at each charge-layer interface to provide relationships between the amplitude coefficients. In the case of bulk-plasmon modes the Bloch ansatz (7.68) still applies, and the same frequencies are obtained as for the infinitely extended superlattice. For the localised surface plasmons, which are characterised by a decaying amplitude with distance from the surface plane z = 0, equation (7.68) holds but with Q replaced by U, where Re(^,) > 0 for the attenuation factor. It is found (see Constantinou and Cottam 1986) that the frequency co of any surface plasmon must satisfy the quadratic equation /*,(//, - 1) smh(qxD)(qxDco)2 + Q[r s (l - 2*0 sinh(qxD) + cosh(qxD)]qxD(o + Q 2 (r 2 - 1) sinh(qxD) = 0
(7.72)
where for simplicity we have ignored retardation effects. The attenuation factor is found from exp( - W) = [rs -
sinh(q
(7.73)
Fig. 7.20 Numerical example of the bulk plasmon dispersion relation (7.71) in the absence of retardation according to the charge sheets model. The plots are in terms of dimensionless variables, co/Q. versus qxD, for three different values of QD: A, 0; B, 7i/2; C, 7i.
280
Plasmons Solutions for co may be written down from (7.72), and they represent surface plasmons only if co is real and |exp( — AZ>)| < 1. When fis = 1 for the surface charge density, (7.72) simplifies to give the solution CD = Q{qxDlcosh(qxD) - rs sinh(qxD)W
- r,2) sinh(qxD)} ^
(7.74)
provided |exp(- XD)\ = \cosh(qxD) - rs sinh(^ x D)|" x < 1
(7.75)
This dispersion relation is consistent with results derived by Giuliani and Quinn (1983). It corresponds to a surface branch above or below the bulk-plasmon continuum if rs < 1 or rs > 1 respectively, and the localisation condition (7.75) implies that there is a cut-off wavevector corresponding to qxD > ln|(l + rs)/(l - rs)|. The more general case of/i s / 1 can be studied using (7.72) and (7.73), and some numerical examples are shown in Fig. 7.21. As fis is reduced from unity it is predicted that each surface branch moves to a lower frequency and the cut-off value for qx is modified. The calculations described so far in this section refer only to intrasubband plasmons, because the excitations involve just one electron band. The 2D
Fig. 7.21 Numerical example of the surface plasmon dispersion relation, calculated from (7.72) and (7.73), for several values of the surface parameters; W, r s = 0.08, fis = 1; X, r s = 0.08, ^ s = 0.75; Y, r s = 4, // s = 1; Z, r s = 4, /xs = 0.75. The shaded region is the bulk plasmon continuum. The other parameters are D = 40 nm and e = 12.9 [after Constantinou and Cottam 1986].
0.5
0
1.0
2.0
3.0
281
Layered structures and superlattices charge sheets model has also been extended to the case of several electron bands, thus providing a description of intersubband plasmons. The formalism becomes much more complicated, and we refer to the work of Tsellis and Quinn (1984) and Eliasson et al. (1987) for details. Other studies have been directed to evaluating Green functions for the superlattice of 2D charge sheets (see, e.g., Jain and Allen 1985a,b,c; Hawrylak et al. 1985; Katayama and Ando 1985; Babiker et al 1986a,b; Eliasson et al. 1987); this has been mainly in the context of light scattering, as we discuss in §7.5.3. The bulk- and surface-plasmon frequencies of charge-sheet superlattices with an alternating structure have also been calculated. For example, Qin et al. (1983) considered superlattices with alternating electron and hole charge layers, while Constantinou and Cottam (1986) investigated the case of superlattices with electron charge layers separated alternately by media A and B that may have different thicknesses and dielectric constants.
7.5.2 Alternating bulk dielectric media A different theoretical model, which allows for finite thickness of the charge layers, is to treat the superlattice as alternating bulk layers, as in §7.3.1 and Fig. 7.10, with an appropriate choice for the dielectric functions s^co) and £2(tt0- For example, we could take sj = fij/>[l- (to?'/co)2]
7=1,2
(7.76)
corresponding to a superlattice in which the two components have charge densities rc, and bulk plasma frequencies < e(b7)
= (nje2/s08^mJ)1/2
(7.77)
and mf are, respectively, the background dielectric constant and Here the effective electronic mass in medium j 0 = 1 , 2). This model may be realised in metallic superlattices or semiconductor superlattices, where the layer thicknesses are sufficiently large that the bulk-like expressions (7.76) for the dielectric function provide a good approximation. In certain cases of interest (e.g. GaAs/A^Ga^^As structures) the electrons may be confined to just one of the media, and we then have the simplification of either c o J ^ O or co{p2) = 0. The excitations are coupled plasmon-polaritons, and their dispersion relations (for s- and p-polarisations) are given by the general results obtained earlier in (7.41)-(7.43). We are principally concerned with low-frequency modes (such that e/o 2 /c 2 « g 2), corresponding to the neglect of retardation effects, because this is the situation for the experiments to be discussed in §7.5.3. In the nonretarded limit there are solutions only 282
Plasmons
for the case of p-modes and the dispersion relation becomes 1 /e, £ 2 \ cos(QD) = c o s h ^ d j cosh(qxd2) + -1 — I — s i n h ^ d j sinh(g 2 \£ 2
ei/
xd2)
(7.78)
This can be alternatively expressed as (e.g. Camley and Mills 1984) -=-jS±O?2-l)1/2
(7.79)
where _ c o s h ^ d j cosh(qxd2) — cos(gD) sinh^Jsinh^^) When ex and £2 have the bulk-plasma response form of (7.76), there are two solutions co± for the frequencies of the coupled plasmon modes: (7.81) where we have introduced the ratios r^co^/ca™ and s = ei^)/ei^). The lower, or 'acoustic', branch with frequency co~ is just the analogue of the superlattice plasmon obtained in the 2D sheets limit. The occurrence of the upper, or 'optic', branch can be associated with finite charge-layer thickness. The dependence of co+ and co~ on qx is illustrated numerically in Fig. 7.22 for GaAs/A^Ga^^As, taking medium 2 to be the charge layers (GaAs) and medium 1 to have no charges (A^Ga^^As), corresponding to r = 0 in (7.81). We have considered afixedsuperlattice period together with various electron-layer thicknesses d2. Other numerical parameters have been assigned in accordance with the Raman measurements of Olego et al. (1982) for their Sample 2, i.e. D = 89 nm, 32 m * = 6.37x 10" kg, e ^ - e j ^ - n . O and no = n2d2 with no = 7.3x 15 2 10 m~ . The experimental sample corresponded to d2/D = 0.29, and we discuss comparison with the Raman data (crosses) in §7.5.3. The above calculations, which were for the dispersion relations of the plasmon-polaritons infinite-thicknesscharge layers, have been generalised to the evaluation of electric-field Green functions by Babiker et al. (1986c, 1987a,b) using the approach discussed in §7.3.2. A limitation of these calculations for the dispersion relations and Green functions is that they employ a scalar dielectric function for the charge layers with the bulk cop value. This would be an unsatisfactory approximation for thin enough layers, and some recent calculations by King-Smith and Inkson (1986, 1987) for this case using a microscopic theory imply that the scalar £ is effectively replaced by a tensor with ezz different from sxx and eyy. Another microscopic theory, also based on RPA, has been put forward by Wasserman and Lee (1985). 283
Layered structures and superlattices 7.5.3 Light scattering
The experimental work on Raman scattering from the plasma oscillations in semiconductor superlattices has been reviewed by Pinczuk (1984). Attention has mostly focussed on the GaAs/A^Ga^^As system with the experiments being carried out under resonance conditions in order to enhance the scattered intensity. The first experiments to demonstrate clearly Raman scattering from plasmons in a GaAs/A^Ga^^As MQW structure were due to Olego et al. (1982). Their light-scattering geometry is sketched in Fig. 7.23, where the incident and scattered light beams are in the same vertical plane (labelled as the xz plane). A right-angle configuration was employed, so 0 + $ = 90° for the angles as measured outside the sample. However, it should be noted that because the refractive index rj of the semiconductor is fairly large (rj ~ 3.6) the incident and scattered beams inside the sample Fig. 7.22 Dispersion curves for the frequencies of the acoustic plasmons (full curves) and the optic plasmons (broken curves) as a function of qxD (for fixed QD). The assumed values of the parameters are given in the text. The corresponding experimental results (Olego et al. 1982) for d2/D = 0.29 are shown (crosses). The theory curves are given for the following values of d2/D: V, 0.0; W, 0.29; Y, 0.6; Z, 1.0 [after Babiker et al. 1986c].
284
Plasmons are near to being antiparallel and parallel, respectively, to the z axis. In this configuration the light-scattering wavevector has components qx = (2ic/X)(sin 9 - cos 0) 2
2
qz = (2n/X)l(rj - sin 0)
1/2
(7.82) 2
2
+ (rj - cos 0)
1/2
]
2
^(2n/X)2rj(l-l/4rj )
(7.83)
where I is the wavelength of light in vacuo and the last line of (7.83) follows using rj»1 (see Problem 7.8). Hence by varying the angle 0, qx may be varied whilst qz is kept effectively constant. Raman spectra due to Olego et al (1982) at three different angles 6 and with X ~ 780 nm are shown in Fig. 7.24. Similar measurements have been reported by Fasol et al. (1985) and Sooryakumar et al (1985). The observed frequencies of the Raman peaks correspond well to the plasmon dispersion relations predicted in §7.5.1 and §7.5.2. This may be seen by reference to Fig. 7.22, where the crosses show data due to Olego et al (1982). Curve V (with d2/D = 0) corresponds to the 2D charge-sheets model and it is seen that theory and experiment are fairly close. However, slightly better agreement is achieved when the finite thickness of the charge layer is taken into account, and the full curve W (with d2/D = 0.29) is appropriate to the experimental sample. The optic plasmon branch for this value of d2/D (see broken curve W) occurs for hco ^ 20 meV. This is close to an intense peak in the Raman spectrum as measured by Fig. 7.23 The geometry for light scattering from plasmons in a superlattice with electron-gas layers.
Incident light
Scattered light
, Dielectric capping layer Electron-gas layers
285
Layered structures and superlattices
Sooryakamur et al. (1985), although those authors attributed it to a coupling between LO phonons and intersubband plasmons. Apart from the peak frequencies, a complete theory of Raman scattering from the plasmons should predict the spectral width and lineshape and the integrated intensity, and also the dependence of these quantities on the scattering geometry and the polarisations of the light beams. Light-scattering calculations have been carried out by several authors (e.g. Hawrylak et al. 1985; Jain and Allen 1985a,b,c; Katayama and Ando 1985) using an approach based on density-density correlation functions or Green functions. In particular, they dealt either with effects in semiinfinite or finite superlattices, where localised surface plasmons are predicted, or with effects of interactions between plasmon and phonons; they found results for the spectral linewidth and shape of the bulk-plasmon and surface-plasmon resonances. A different approach, based on general methods for calculating light scattering from polaritons in superlattices, was adopted by Babiker et al (1986a,b,c; 1987b) for both the ID sheets Fig. 7.24 Raman spectra for scattering from plasmons in a GaAs/Al^Ga^^As MQW structure for three different angles 6 in a right-angle scattering geometry [after Olego etal. 1982].
1
3 Stokes shift (meV)
286
5
Magnetic properties and thefinite-thicknesscharge-layer models. The formalism incorporated the dependence on the scattering geometry and the polarisation directions. For a fuller understanding of light scattering from plasmons, it would be helpful to have more experimental data, particularly for the integrated intensities and for systems other than GaAs/Al^Ga^^s. More work, both experimental and theoretical, should be directed towards the superlattice surface plasmon. 7.6 Magnetic properties We finally give an introductory account of superlattices in which one or both of the component materials are magnetic. The nature of the magnon excitations depends crucially on whether the long-range dipoledipole interactions or the short-range exchange interactions dominate, just as discussed in Chapters 3 and 4 for single magnetic layers, and this in turn depends largely on the excitation wavevector. We treat the case where dipolar terms are dominant (i.e. the magnetos tatic and magnetic polariton regimes) in §7.6.1, and then deal with the exchange coupling in §§7.6.2 and 7.6.3. It is towards the former case that most experiments to date have been directed, with the main results coming from light scattering and microwave absorption. 7.6.1 Magnetostatic regime and magnetic polaritons The simplest case is a superlattice in which ferromagnetic and nonmagnetic layers alternate, as represented in Fig. 7.25. The magnetostatic modes for this case can be calculated by a direct extension of the theory for a double-layer ferromagnet (see §4.3) and so we shall quote only the final results. The usual geometry corresponds to the applied field and magnetisation directions being parallel to the interfaces as in Fig. 7.25. The theory was first worked out by Camley et al. (1983b) and Griinberg and Mika (1983). The former authors calculated the magnetic Green functions as well as the dispersion relations, while the latter authors concentrated on the dispersion relations but included effects of a finite number of layers in the superlattice. For the special case of the in-plane wavevector q^ perpendicular to M o (the Voigt configuration), the results for the magnon frequencies may be summarised as follows: the frequencies can be conveniently expressed in terms of a parameter a, as in (4.37) for magnetic double layers, where the modes of the superlattice correspond to: (i) a band of bulk modes for which 2sinh(^||rf1)sinh((?l|rf2) cosh(4,|D)-cos(gxD) where q = (qx9 qy) is the 3D wavevector, and
(
}
287
Layered structures and superlattices (ii) a surface-magnon branch that has (for a semi-infinite superlattice) a=0
(7.85)
and exists only if d2>d1. As a numerical example we show in Fig. 7.26 the values of a for bulk and surface modes of the superlattice plotted against d1/d2 for fixed values of d2 and q^. It can be shown that the bulk-magnon mode of the superlattice is made up of a linear superposition of surface waves within each magnetic layer; the net effect is a bulk mode since the amplitudes and phase factors are such that Bloch's theorem is satisfied (see Camley et al. 1983b). The bulk modes always lie in the frequency range from [(O0(co0 + co m )] 1/2 to (co0 + icom), in the notation of Chapter 4. The surface mode of the superlattice is also a linear superposition of surface waves within each magnetic layer, but they are combined with an envelope function that decays exponentially with distance into the superlattice. This mode, which has the property of existing only if the magnetic-layer thickness d2 is greater than the spacer-layer thickness dl9 has frequency (co0 + jcom) as can be deduced from (4.37) and (7.85). This frequency is
Fig. 7.25 The assumed geometry for a semi-infinite magnetic superlattice composed of alternating ferromagnetic and nonmagnetic layers.
Ferromagnetic layer Nonmagnetic spacer
288
Magnetic properties
the same as that of the surface mode in a semi-infinite ferromagnet (see §4.2). Brillouin spectroscopy has been used to test experimentally the above prediction for the surface mode of the superlattice (Grimsditch et al. 1983; Kueny et al. 1984). In Fig. 121(a) and (b) we show examples of spectra obtained for Ni/Mo superlattices with d2 > dl and d2
Bulk modes of superlattice
Surface mode of superlattice -0.2
Fig. 7.27 Brillouin spectra for Ni/Mo superlattices in an applied magnetic field of 0.093 T: (a) d2 = 24.9 nm, d1 = 8.3 nm; {b) d2 = 10 nm, dx = 30 nm [after Grimsditch et al. 1983] 4 2 (a) (b)
I 0 0.4 -0.4 Frequency shift (cm"
! -0.4
0.4
289
Layered structures and superlattices
confirmation of the prediction concerning the surface magnon of the superlattice. Magnetostatic theory has also been applied to alternating ferromagnetic/ nonmagnetic superlattices for the case of the magnetisation Mo perpendicular to the interfaces (Camley and Cottam 1987). We already remarked in §4.2 that, for a single ferromagnetic slab in this perpendicular magnetisation case, no surface magnetostatic modes occur. However, the above authors showed that, when a superlattice of these magnetic layers is formed, the coupling between layers is such as to allow a sequence of surface magnetic modes in appropriate cases. Calculations were presented for both ferromagnets and antiferromagnets. A few theoretical papers have appeared generalising the above results to include retardation. Barnas (1987) deals with ferromagnetic/nonmagnetic superlattices, while a later paper (Barnas 1988) uses the transfer-matrix method to derive general formal results for both bulk and surface modes in magnetostatic, polariton and exchange-dominated regimes. The polariton regime was also discussed by Raj and Tilley (1987). All the results derived are for the Voigt configuration. As might be expected, the full results are quite complicated, but fortunately they can be greatly simplified in the long-wavelength limit. As shown by Raj and Tilley, the method due to Agranovich and Kravtsov (1985) can be applied to a magnetic superlattice as follows. We consider essentially the geometry of Fig. 7.25, except that both constituents are now taken to be magnetic. For propagation in the xy plane (Voigt geometry) each constituent is characterised by a gyrotropic permeability tensor of the form of (6.102). We consider an rf magnetic field in the xy plane; the wavelength is taken to be much longer than the superlattice period so that the B and H fields are the same in successive layers of component 1 and in successive layers of component 2. The boundary conditions are that Hy and Bx are continuous across interfaces, so these components are everywhere equal to their average values Hy and Bx. In medium 1 the field components are related by (7.86)
(7.SI) so that B ^
= fii'% + (fi^/n^BJfio x
Hx = bixyii™)Ry + (l//JLx i)EJii0
(7.88) (7.89)
where /xv = fixx + \i\yl\ixx is the Voigt permeability. Similar equations hold in component 2, and combined with (7.88) and (7.89) give expressions 290
Magnetic properties + d2B(y2))/(d1 + d2) and Hx:
for the average fields By = (d^
(7.90) (dx + d2)Hx = (dii/^/Hx1* + d2fiixy/^ix2i)Ry + (di/n^x + dJfi^BJfiQ Reorganisation gives relations in terms of the effective permeability tensor: /***
(7.91) medium
fixy\f
- fixy fiyyj \Hyy
with " -
^yy
/i
(7.93)
.
i
w
i
r?\
.
i
nu
('•"^•;
(7 95)
-
This derivation was given independently by Almeida and Mills (1988). The form of (7.92) corresponds to a kind of magnetic anisotropy, induced by the superlattice structure itself. It is straightforward to find the dispersion equation for wave propagation in a medium with this kind of magnetic anisotropy. For a TE mode, the relevant component of the dielectric tensor in the effective-medium description is syy: syy = (e1d1+e2d2)/(d1+d2)
(7.96)
This dispersion equation is Vyy + VxylVxx
Vxx + PxylPyy
Raj and Tilley (1987) show that this is the same form as can be obtained by a long-wavelength expansion of the general dispersion equation, derived by the transfer-matrix formalism. 7.6.2 Exchange-dominated properties: reconstruction Some exploratory theoretical work has been carried out for situations in which the exchange interaction is dominant and dipole-dipole effects can be neglected. It is important to recognise at the outset that even the nature of the ground state depends on the specific details of the system under consideration, and that in the presence of external fields and with variation of temperature many forms of spin reconstruction are possible. This is nicely illustrated by the hypothetical ferromagnetic/ antiferromagnetic superlattice discussed by Hinchey and Mills (1986a,b). 291
Layered structures and superlattices The system consists of alternating ferromagnetic and antiferromagnetic components, with a ferromagnetic coupling at the interface. The antiferromagnetic structure is one of alternating layers, successively spin-up and spin-down. It can be seen that the nature of the ground state depends crucially on whether the number NAF of layers in the antiferromagnet is even or odd. This is illustrated in Fig. 7.28. For NAF even, Fig. 7.28(#), successive ferromagnetic layers are aligned oppositely, and the ground state has no net magnetic moment, while for NAF odd, Fig. 7.28(c), the ferromagnetic layers are all aligned, and the ground state does have a magnetic moment. Even at zero temperature, the phase diagram in an applied external field of the system in Fig. 7.28 is quite complicated. Figure 1.2$(b) shows a possible state of the NAF-even superlattice in which all the ferromagnetic spins are aligned with the applied field. The Zeeman energy is lower than in the configuration of 7.28(a), but the interface-exchange energy is higher. As the number of spin layers Ns = NF + NAF in one period increases the relative importance of the interface-exchange energy decreases, so it is to be expected that the critical field Bc for a phase transition out of the state of Fig. 1.2$(a) decreases as Ns increases. Hinchey and Mills find Bc by a 'soft-mode' method, that is, they find Bc as the field for which a spin-wave Fig. 7.28 Ferromagnetic/antiferromagnetic superlattice. (a) Zero-field ground state when number of antiferromagnetic layers NAF is even, (b) Possible high-field state, (c) zero-field ground state when NAF is odd [after Hinchey and Mills 1986].
it 11 tit i i lit m i l t i ttj -j - «!
J
!
^
!
dx
*t-
d2
It mint i 292
!
AF
i
+~—
:
F
dY
i
i
AF
i
I
Magnetic properties frequency becomes zero. Typical results are shown in Fig. 7.29; the field units are such that the spin-flop transition of the bulk antiferromagnet corresponds to Bo = 1. As expected, for even N AF the zero-field ground state is rather unstable, with Bc -+0 as N -> oo, while for odd NAF the zero-field ground state is rather stable. The transition out of the ground state at the critical field Bc in Fig. 7.29(a) is not straight into the 'Zeeman state' of Fig. 7.28(ft); in fact what occurs is a 'twisted state' in which the spins gain a y component, so that they remain in the plane of the layers but are no longer aligned with the magnetic field. A detailed discussion of the twisted states and phase diagrams is given by Hinchey and Mills. Spin reconstruction is also discussed by Camley (1987) and Camley and Tilley (1988). They are concerned with the Fe/Gd system, which has Fig. 7.29 Critical fields Bc for a transition out of the ground state of the ferromagnetic/ antiferromagnetic superlattice as functions of N = iVp. (a) NAF even, corresponding to Fig. 7.28(a), (b) JVAF odd,corresponding to Fig. 7.28(6). The units of field are chosen so that the bulk AF spin-flop field is equal to unity [after Hinchey and Mills 1986]. 0.20
0.15
0.10
0.05
0.00
H
1
h
1.50
1.00
Bulk AF spin-flop field
0.50
0.00
10
12
14
293
Layered structures and superlattices
the interesting property, as demonstrated by Taborelli et al. (1986), that although Fe and Gd are ferromagnets, the exchange interaction at an interface is antiferromagnetic. The ground state of a Fe/Gd superlattice is therefore as depicted in Fig. 7.30(a). This may be expected to be unstable in modest applied fields, and indeed Camley and Tilley find that a phase transition occurs to a twisted state as shown in Fig. 7.30(b). Numerical exploration of the mean-field expression for the free energy can be used to find the phase diagram in the B0T plane. The detailed results, presented by Camley and Tilley (1988), show that for small values of n1 and n2 the phase diagram is sensitive to the precise number of spins in each layer. A formal development that should be of value in discussing reconstruction is the use of Landau-type expansions, as proposed by Fishman et al. (1987) and by Camley and Tilley (1987). The method applies only for the case when nx and n2 are large, and when the average spin value varies only slowly from site to site. The mean spin <Sf> at site i may then be replaced by a continuous variable M(x). The basic idea of Landau theory, as expounded in Landau and Lifshitz (1969) for example, is to expand the free energy as a Taylor series in the invariants of the system.
Fig. 7.30 Magnetic superlattice of Fe/Gd. (a) Ground state in absence of magnetic field, {b) Twisted state developed in applied field Bo [after Camley and Tilley 1988].
layers Fe
layers Gd
(a)
294
Magnetic properties For a slab of Fe, extending from x= —d1 to x — 0, the expression is 1==
f00
f°
dydz
J - oo
J -di
,
]}
(7.98)
Here a l 5 b x and c1 are constants, which may be related to the microscopic parameters for Fe, or may be determined later from comparison with experiment. It is assumed that M is confined to the yz plane, corresponding to the twisted state of Fig. 7.30(b). A similar expression for a term F2 in terms of a variable N, say, holds for the other component of the unit cell, namely a slab of Gd extending from x = 0 to x = d2. Finally it is necessary to include a free energy term for the interface. The lowest order invariants are (M 2 + M 2 ) x = 0 , (AT2 + N2)x = 0 and (MyNy + MZNZ)X=O, so that the interface free energy is
-I"
•N2)x
= 0
(7.99)
J -c
This introduces three more parameters 8l9 b2 and a. Within this model, the equilibrium state is now found as that which minimises F1 + F2 + FY. The minimum is found, in principle, from the Euler-Lagrange equations for M and N. However, these are quite complicated, and we simplify the problem by use of the constant-amplitude approximation: (My, Mz) = M(sin 6, cos 0)
(7.100)
(Ny, Nz) = N(sin 0, cos >)
(7.101)
This approximation should be good at low temperatures, where the reconstruction is described in terms of spins of constant magnitude inclined at angles 0 and (j) to the field direction. Within this approximation, the free energy reduces to — B0M cos 0 + \c^
f
*<)
Jo + <xMN cos(0 - 4>)\x=o + aMN cos(0 -
(7.102)
Here certain constant terms have been omitted, and it is assumed that the equilibrium state has period d1+d2, so that (7.102) contains contributions from the two interfaces in the unit cell. A is the specimen area in the yz plane. 295
Layered structures and superlattices The Euler-Lagrange equations for minimum F are c xM — - - B dx 2
o
sin 0 = 0
(7.103)
and c2N —^ - B o sin 0 = 0 dx with boundary conditions at interfaces d6 c,M — -OLN sin(0 ->) = 0 dx
(7.104)
(7.105)
c2AT-^ - aM sin(0 - 0) = 0 (7.106) dx In general the solutions of (7.103) to (7.106) may be given in terms of elliptic functions, as shown by Camley and Tilley (1988). For the particular case of two semi-infinite media in contact, the elliptic functions reduce to hyperbolic functions, and the solutions are cos \Q = tanhCA^j^! - y)] cos \4> = tanh[2 2 (j/ - y2)]
(7.107) (7.108)
where X\ = BolClM
(7.109)
with a similar expression for X\. yx and y2 are constants of integration, found from the boundary conditions, or equivalently from minimisation of the free energy. The details are left for Problem 7.10, and the predicted reconstruction for the case of two semi-infinite media in contact is shown in Fig. 7.31. It can be seen that, as might be expected, the width of the reconstructed region decreases as the applied field Bo increases. The Landau theory described here should in principle be applicable to ferroelectric superlattices, which are conceivable, although to date no examples have been reported. 7.6.3 Exchange-dominated properties: magnons In principle the magnon spectrum must be derived for each of the equilibrium states found for a particular superlattice, and in fact Hinchey and Mills (1986a,b) do give an extensive discussion for the ferromagnet/antiferromagnet superlattices with which they are concerned. The dispersion relation can be found by application of the transfer-matrix method to spin equations of motion, as used in Chapter 3. We give here an illustration for what is probably the simplest case, the ferromagnet/ ferromagnet superlattice with ferromagnetic exchange at the interfaces, so 296
Magnetic properties
that reconstruction need not concern us. This has been discussed by Albuquerque et al. (1986b) and by Dobrzynski et al. (1986). To be specific, following Albuquerque et al. (1986b), we consider alternating simple-cubic ferromagnets in which the interfaces are (001) planes. It is assumed that each component is described by the nearestneighbour Heisenberg Hamiltonian including an external field, (3.4), and the exchange constants are taken as Jt in one component, J2 in the other, and / across the interface. The structure is shown in Fig. 7.32. The problem is essentially one-dimensional, and the transfer matrix may be constructed exactly as described in §§7.1.1 and 7.2.1. Spin equations of motion for the operators S* are written down, and simplified by using RPA (see §3.2). Within each component the solution for S* is written as a sum of forward- and backward-travelling waves, as in (7.3). The equations of motion of spins adjacent to the interfaces, labelled a, /?, y, S in Fig. 7.32, relate amplitudes in adjacent layers, and the transfer matrix is built up Fig. 7.31 Variation of 6 and (j> with position x for a single interface between Fe and Gd. Magnetic field varies from Bo = 0.2 T (broken line) to 1 T (chain line) in steps of 0.2 T. Horizontal axis scaled by AO = 5.85 x 10 7 m~ 1 [after Camley and Tilley 1988].
10.0
-120
297
Layered structures and superlattices as in §7.2.1. Finally, the dispersion equation takes the standard form of (7.23). An example of the dispersion curves obtained for this model is shown in Fig. 7.33. Here the bulk-medium dispersion curves of Fig. 7.33(a) are drawn from (3.12), and the Brillouin-zone edge frequencies are denoted co1 and co2- The resemblance to Fig. 7.7 for diatomic-lattice phonons is striking. For a> < CD1 the bulk wavevector q1 and q2 in the two components are real, and the superlattice dispersion curve consists of broad pass bands with narrow stop bands. For (O1
II
o o o o o o o o o o O
O
O
y $
IJ
Or O/
o o o o o o o o o o n, a > < a > << (l-l)na
o
o
o
o
o
o
o o o o
Ina
Fig. 7.33 Bulk and superlattice magnon dispersion graphs for (100) propagation with parameters g^iBB0/4SJl = 1, J2/J1 = 2,1/J^ = 1.5. Both components are taken to have the same values of gyro magnetic ratio g and spin S. Relevant frequencies are hcoJgnBBo = 1 + 4SJJgnBB0 = 2, hco2/gnBB0 = 1 + 4SJ2/gnBB0 = 3. (a) Bulkmagnon dispersion curves for components 1 (lower curve) and 2 (upper curve), (b) Superlattice magnon dispersion curves for n1=n2= 10 [after Albuquerque et al. 1986b]. 0)9 = 3 r
&3
CO
I"
qta
298
Problems the superlattice dispersion curve then consists of narrow pass bands and broad stop bands. Barnas (1987b) generalises the formal results for a ferromagnet/ ferromagnet superlattice in a number of ways. First, he considers the case when the repeat unit consists of N different ferromagnetic layers; second, he includes uniaxial anisotropy; third, he discusses the surface magnons of a semi-infinite specimen as well as the bulk magnons. Particularly in view of the wide variety of possible combinations of ferromagnetic and antiferromagnetic exchange it is clear that the theoretical study of reconstruction and exchange-dominated magnons is still at an early stage. Likewise, at the time of writing most of the experimental work remains to be done. 7.7 Problems 7.1 In §7.4, the lattice dynamic calculations are given for a diatomic superlattice. Perform the equivalent calculations for a monatomic superlattice. Check that appropriate limits are found, in particular that the continuumacoustics results of §7.3 are found as the long-wavelength limit of your results. 7.2 Equations (7.39) and (7.40) give the diagonal elements 7\ , and T22 of the transfer matrix T for s-polarisation. Evaluate T completely, and confirm that T is unimodular, det T = 1. 7.3 Evaluate the transfer matrix T for p-polarised optical propagation. Confirm that T is unimodular, and that the dispersion relation is given by (7.41) and (7.43). 7.4 Show that the transfer matrix for the electron amplitudes in (7.55), defined by
\bl + l)
\bt
has elements
T2i=kfxy(s-l-s) where A
= =
t2-t1
299
Layered structures and superlattices tj =ik1/m1
t2 = ik2/m2
f=exp(ikldl) s = Qxp(ik2d2) Hence prove that T is unimodular, det T = 1, and derive the dispersion relation (7.62). 7.5 Starting from the expressions (7.65) and (7.67) for the electric and magnetic fields within the dielectric media of a 2D charge-sheets superlattice, prove the dispersion relation quoted in (7.69). You should apply the boundary conditions that
at the interface z = nD, as well as Bloch's theorem in the form (7.68). Here a is the 2D conductivity given by a = in0e2/m*co
in the notation of §7.5.1. From the solvability condition for the simultaneous equations relating coefficients ain) and b(n) you should obtain (7.69). 7.6 Prove that for small enough qx (such that qxD « 1) the dispersion relation (7.71) implies that w is proportional to qy (except when cos(QD)= 1). Find the constant of proportionality, and investigate the case where cos(QD)=\. What is the approximate dependence of co on qx for the opposite limit of qxD»V. 1.1 Rearrange (7.78) to verify the alternative form of the dispersion relation for nonretarded p-modes given by (7.79) and (7.80). Hence obtain the dispersion relation for acoustic and optic plasmons given by (7.81). 7.8 Prove the last step in (7.83) for the light-scattering wavevector by approximating the square roots using rj» 1. 7.9 Confirm that within the constant-amplitude approximation the Landau free energy Fx + F2 + F, reduces to (7.102), and that the conditions for a minimum in (7.102) are those given by (7.103) to (7.106). 7.10 This is concerned with the solution of the constant-amplitude equations (7.103) to (7.106) for two semi-infinite media in contact, (a) Prove that (7.103) has the first integral
and note that the boundary conditions 0-»O as x —> — oo 300
and
Problems d0/dx->O as x-> -oo imply that l\ = X\. (This simplification does not occur for the superlattice.) (b) Hence confirm (7.107) and (7.108) for 9 and >. (c) Prove that upon substitution of (7.107) and (7.108) the free energy (7.102) may be written F-FA = (4BOM/A1)(1 -tt) + (4BONM2)(1 - t2) + *MN{(2t\ - l)(2t| - 1) - 4 ^ ( 1 - t\)l'2(\ - ti)1'2} where r2 = tanh(-/l2>;2) and FA is the free energy of a hypothetical state with 9 = § = 0 and a = 0. Minimisation of F with respect to tx and t2 determines yx and y2, and hence enables the curves of Fig. 7.31 to be drawn.
301
8 Concluding remarks
In the preceding chapters we have presented a survey of the dynamical properties of planar surfaces and interfaces. The selection of surface excitations, with which we illustrated the general principles, was wide ranging but inevitably incomplete. Our choice was guided by the existence of appropriate theoretical models, by the availability of data to compare theory and experiment, and in particular by recent advances in fabricating low-dimensional solid structures. Several topics relevant to the study of surface excitations have been omitted, or mentioned only in passing, in the previous chapters. It is useful now to provide some brief comments on these topics, together with references to one or two examples in each case. 8.1 Mixed excitations Apart from polaritons, which are excitations formed when a crystal excitation couples to a photon, and which were extensively discussed in Chapter 6, we have not considered 'mixed' modes of two excitations. These may be important for appropriate frequency and wavevector values, and as a relatively straightforward example we take the case of coupled magnetic and vibrational modes. In the long-wavelength limit, where the solid can be treated as an elastic continuum, these modes are known as magnetoelastic waves. For a ferromagnetic solid with simplecubic lattice structure and static magnetisation in the z direction, the energy of the magnetoelastic coupling for the magnon regime where only terms linear in the magnetisation M are retained takes the form (Kittel 1958) b
[Mx(uxz + uzx) + My(uyz + uzy)2 Mo' 302
(8.1)
Nonlinear effects and interactions between excitations Here b is a phenomenological coupling constant, Mo is the static magnetisation, and the u^ are elements of the strain tensor (see §2.2.1). In the absence of magnetoelastic coupling (b = 0) the excitations of a semi-infinite medium consist of (i) the bulk and surface elastic waves obtained from the equation of motion (2.36) with appropriate boundary conditions as discussed in §§2.2 and 2.3, and (ii) the bulk and surface magnons obtained from the torque equation (3.35) or (4.11) as discussed in §§3.3 and 4.2, respectively, for different magnetic models. When b ^ 0 the equations of motion must be supplemented by additional terms due to the coupling (8.1). This has been discussed, for example, by Scott and Mills (1977). These authors concentrated on a theoretical study of surface acoustic-type waves propagating in a semi-infinite ferromagnetic medium, taking the long-wavelength regime where exchange effects can be ignored (i.e. the magnetostatic limit). In particular, they presented calculations for two types of surface acoustic wave, one being in essence a Rayleigh wave (see §2.3.1) modified by the magnetoelastic coupling and the other a shear-like magnetoelastic surface mode that exists only because of the presence of the magnetoelastic coupling. These waves have nonreciprocal propagation characteristics, a consequence of their being formed by coupling to the Damon-Eshbach surface magnetostatic mode which is nonreciprocal as was discussed in §4.2. Scott and Mills conclude that these magnetoelastic surface waves can provide a useful and sensitive probe of the bulk and surface magnetostatic modes, especially at larger values of the in-plane wavevector qy. Other references to magnetoelastic surface waves in ferromagnets are to be found in the review articles by Maradudin (1981) and Cottam and Maradudin (1984). Calculations have also been performed for Rayleigh waves in paramagnetic rare-earth systems (Lingner and Luthi 1981; Camley and Fulde 1981). In this case the magnetoelastic interactions describe the coupling of the strain and the rotational part of the deformation tensor to the unfilled 4f electron shell. Paramagnetic crystals are of interest because their elastic moduli typically exhibit a strong dependence on temperature and applied magnetic field, leading in turn to strong temperature and magnetic field dependences of the Rayleigh wave propagation. Experiments to measure the Rayleigh wave velocity have been reported for CeAl2 and SmSb (Lingner and Liithi 1981), and are in good agreement with the theoretical predictions. 8.2 Nonlinear effects and interactions between excitations In practice all the excitations discussed in this book may exhibit nonlinearities to some degree. The nonlinearities lead to interactions between the otherwise independent solutions (normal modes) of the 303
Concluding remarks equations of motion for the system. Important consequences are that the mode frequencies become modified (or 'renormalised') due to the interactions and also the modes are 'damped', i.e. they acquire a finite (rather than infinite) lifetime. Other nonlinear effects include second (and higher) harmonic generation of surface and bulk excitations and nonlinear mixing of two (or more) excitations. An example of nonlinearity involving electromagnetic waves has already been discussed in §6.4, and we now give some additional examples involving vibrational and magnetic excitations. In the case of vibrational modes in semi-infinite elastic media, there have been extensive studies of second-harmonic generation and parametric mixing of Rayleigh surface waves. According to nonlinear elasticity theory (see Wallace 1970), there are forces created within the medium of the form
NL
c
/* = I ^^tlt
(8 2)
-
and surface stresses of the form 7^NL_
y [ivpa
(S)
<^<^p CXV
(8 -x\
OXa
In (8.2) and (8.3) the quantities c(b) and c(s) are combinations of third order elastic moduli; {u^} and {x^} denote Cartesian components of the displacement vector and position vector respectively. It is the occurrence of nonlinearities in the boundary conditions (resulting from (8.3)) as well as in the equations of motion that makes the analysis particularly complicated. The first theoretical work on second-harmonic generation of Rayleigh waves was reported by Viktorev (1963), and the first experiments were due to Rischbieter (1965, 1967) for Al and steel surfaces. Later work has been reviewed by Maradudin (1981) and Stegeman and Nizzoli (1984). As a further example we mention some calculations for semi-infinite Fig. 8.1 The four-magnon interaction dominant at low temperatures T
304
Excitations in wedges and at edges
Heisenberg ferromagnets where higher-order magnon-magnon interactions have been taken into account by using better approximations than the simple RPA discussed in Chapter 3. At low temperatures T«T C the dominant interaction is a four-magnon process in which two incoming magnons scatter to produce two outgoing magnons (see Fig. 8.1). The magnons may be surface or bulk modes in any combination. The damping (i.e. /?/T, where T is the lifetime) of surface magnons was calculated by Tarasenko and Kharitonov (1974). They found that in the absence of an applied magneticfieldand anisotropyfield,the damping of long-wavelength surface magnons is approximately proportional to q\T2 at T« Tc. This contribution comes from interactions of the surface magnon with three other surface magnons, and the result takes this form provided T/Tc« a2q\ « 1, where a denotes the lattice constant. Calculations for the renormalisation of the surface-magnon frequencies due to magnon-magnon interactions have been given by Kontos and Cottam (1984, 1986) and Mazur and Mills (1984); these results depend sensitively on the lattice structure and on the surface values of the exchange and anisotropy parameters. 8.3 Excitations in wedges and at edges
An interesting generalisation of the parallel-sided slab geometry, which we have discussed in many applications, is to the case where two planar surfaces meet at an angle 0 to form a wedge-shaped sample. The surfaces are now such as to destroy translational symmetry in two different directions, with obvious consequences for the symmetry of the excitations. In the geometry of Fig. 8.2 the apex of the wedge corresponds to the line x = y = 0. There is a translational invariance along the z direction Fig. 8.2 A wedge of angle 6 showing the assumed orientation of coordinate axes.
305
Concluding remarks
(assuming the wedge dimensions to be effectively infinite in this direction) but not along the x or y directions. This means that, under appropriate conditions, there may be localised excitations known as wedge modes that propagate in a wave-like fashion along the apex of the wedge but decay with increasing distance into the wedge from its apex and faces. The case of a right-angled wedge (0 = 90°) often leads to simplifications, particularly in microscopic calculations involving cubic materials, and we refer to the localised excitations in this situation as edge modes. However, the terms 'wedge modes' and 'edge modes' are often used interchangeably in the literature. The first calculations were for acoustic waves in wedges and edges formed by isotropic, elastic media (Lagasse 1972; Maradudin et al. 1972). The wedge modes have certain convenient properties, such as being essentially nondispersive, which make them of interest for waveguide applications. The number of wedge modes, and their speed of propagation, can be controlled by varying the wedge angle 0. In general, as 0 is decreased the number of modes increases and their propagation speeds become lower than that of Rayleigh waves. This is illustrated by a numerical example in Fig. 8.3. Apart from the ideal infinite wedge, calculations have been Fig. 8.3 The square of the speed of propagation of wedge modes (in units of the speed of bulk transverse acoustic waves) of T 2 symmetry as a function of wedge angle 6. The speed corresponding to Rayleigh waves is denotes by QR [after Moss et al. 1973].
306
Excitations in spheroidal and cylindrical samples
applied to more realistic waveguide structures, e.g. where the wedge-tip is rounded rather than sharp, or where the waveguide consists of a wedgeshape ridge on a substrate material. A thorough review has been given by Maradudin (1981). Electromagnetic wedge modes have also attracted much theoretical attention (see Maradudin 1981). Most calculations have employed the electrostatic approximation (see §6.2.2) since this considerably simplifies the analysis. In an infinite dielectric wedge surrounded by vacuum, one may then solve Laplace's equation by the method of separation of variables to obtain a scalar potential proportional to exp(igz) and localised about the apex of the wedge. Details are given by Dobrzynski and Maradudin (1972). When retardation effects are included, one is faced with the much harder task of solving the full set of Maxwell's equations in the wedge geometry; an elegant calculation of retarded wedge modes in a parabolic wedge has been given by Boardman et al. (1981). The magnons localised at a 90° edge of a Heisenberg ferromagnet with simple-cubic structure have been studied by Sharon and Maradudin (1973) using a long-wavelength continuum method and extended to shorter wavelengths by Maradudin et al. (1977) using a microscopic theory. They used a model in which all exchange interactions (including those involving spins at the surface) have their bulk values, denoted by J for nearest neighbour and j for next-nearest neighbours. At long wavelengths and in the absence of surface anisotropy fields, the magnon amplitudes of the edge modes have the approximate dependence exp[ - K(X + yj] exp[i(qz - cor)] (8.4) where the attenuation constant K is related to the propagation wavevector q by (8.5) provided K > 0. This is an acoustic-type branch occurring split off below the surface magnon, which is itself split off below the bulk continuum. This behaviour is illustrated in Fig. 8.4. 8.4 Excitations in spheroidal and cylindrical samples
Throughout this book we have idealised the surfaces and interfaces as being planar and smooth. Some generalisations to curved samples (in particular, spheroids and cylinders) are now presented, and the topic of surface roughness is treated in §8.5. We start by considering magnetostatic modes in ferromagnetic samples with curved surfaces. The case of spherical and spheroidal samples was 307
Concluding remarks worked out by Mercereau and Feynman (1956) and Walker (1957), followed a few years later by the case of a cylinder (Fletcher and Kittel 1960; Joseph and Schlomann 1961). The results are reviewed by Wolfram and Dewames (1972). We briefly consider the calculation for a long circular cylinder magnetised along its axis of symmetry (the z direction). A scalar potential i// may be introduced as in §4.2, and this must satisfy (4.21) and (4.22) inside and outside the sample respectively. After transforming to cylindrical coordinates (r, 9, z) in the usual way, one seeks separable solutions of the form iA(r, 0, z) = p(r) exppfoz + m0)]
(8.6)
Here m is a positive or negative integer, qz is the propagation wavevector in the z direction, and the radial part p(r) of the potential has to satisfy q2zp = 0
(8.7)
in the interior and a similar equation with xa = 0 in the exterior. The acceptable solutions of (8.7), which are regular at r = 0 and vanish as r -• oo, can be written down in terms of Bessel functions. It then remains Fig. 8.4 Calculation of dispersion relations for bulk magnons (full line), surface magnons (short dash) and edge magnons (long dash) in a simple-cubic Heisenberg ferromagnet with a 90° edge. The bulk dispersion curve refers to the lower edge of the continuum [after Maradudin et al. 1977]. 12.0 -
J'/J = 0 . 5
9.0
(D/JS
6.0
3.0
0.0
308
0.0
0.5
1.0
Surface roughness
to apply the boundary conditions at r = R, where R is the radius of the cylinder; these take the form of continuity of \j/ and continuity of the radial component of the magnetic flux density b at r = R. The resulting dispersion relation has a sequence of solutions, characterised by the positive index p (p = 1, 2, 3 , . . .), for the surface modes. In a long-wavelength limit (qzR« 1) the frequencies simplify to p=\ (8 8) p>i ' These surface magnons all have frequencies above those of the bulk manifold, i.e. cos > [coo(a)o + com)]1/2. I*1 general, the localisation at the surface of the cylinder increases with increasing qz; details are to be found in the original references. The magnetostatic calculation for a spherical or spheroidal ferromagnet can be carried out by an analogous generalisation of §4.2; the solutions for the potential ij/ involve Legendre functions instead of the Bessel functions of the cylindrical case. The theory of elastic surface waves in homogeneous isotropic materials, as discussed in §2.3 for a parallel-sided slab geometry, can be extended relatively straightforwardly to cylindrical and spherical geometries. The propagation of Rayleigh-type surface modes can then be studied in appropriate cases. The first calculation was due to Hudson (1943) for elastic waves travelling along the axis of a long circular cylinder; the considerable amount of subsequent work has been reviewed by Maradudin (1981). Another example is provided by surface and bulk polaritons at curved surfaces. The polariton dispersion relations and the associated electromagnetfielddistributions have been calculated for spheres and for circular cylinders using similar methods to those in Chapter 6. For example, phonon-polaritons in the absence of retardation have been considered by Englman and Ruppin (1968a); the theory was subsequently generalised by the same authors to include retardation (Ruppin and Englman 1968; Englman and Ruppin 1968b). Further references are contained in the review articles by Economou and Ngai (1974) and Kliewer and Fuchs (1974).
8.5 Surface roughness
All real solid surfaces are rough to some degree. This may be because of the way in which the sample has been prepared, e.g. by vapour deposition on to a substrate material, or because of subsequent treatment of the surface, e.g. polishing of the surface by an abrasive substance of a particular particle size. It is therefore relevant to determine how this roughness affects the properties of the excitations and how these properties 309
Concluding remarks
differ from those in samples an an 'ideal' plane surface. The use of a rough surface to couple external radiation out of the surface plasmonpolariton generated in a tunnel junction was mentioned in §6.6.2. In many cases the presence of surface roughness acts simply as a small perturbation on effects that already occur with plane surfaces, e.g. the frequency and/or damping of an excitation may be perturbed due to the roughness. For example, the modifications of Rayleigh surface waves propagating along a randomly rough surface have been studied by Urazakov and Falkovsii (1972) and by Maradudin and coworkers (see Maradudin 1981). The latter calculations were carried out by a perturbation method correct to order (d/b)2, where d is the root-mean-square departure from flatness and b is the mean distance between consecutive peaks and valleys on the surface. In other cases the random roughness can produce qualitative changes; an example is the roughness-induced splitting of the surface-plasmon dispersion curve observed experimentally (e.g. Palmer and Schnatterly 1971) and studied theoretically (e.g. Kretschmann et al. 1979). It can also produce large quantitative effects, such as the giant enhancement (by five orders of magnitude or more) of the Raman intensity for scattering from molecules adsorbed on certain metal surfaces. Although there are various mechanisms proposed for this enhancement, it is claimed that surface roughness can give rise to a significant part of it (Burstein et al. 1982). This is supported by the observation by Ushioda et al. (1979) that the Raman intensity for scattering from surface polaritons in GaP increased by a factor of four tofivewhen the surface of the sample was roughened. Apart from the random roughness referred to in the preceding paragraphs, there can be periodically corrugated surfaces, e.g. as in the case of a diffraction grating ruled on the surface. The use of a grating to couple external light to a surface polariton was explained briefly in §6.6.1. Indeed, grating surfaces are of considerable technological interest because of their role in optical surface acoustic wave devices, for example as surface mode to bulk wave transducers. Also, from the theoretical point of view, they allow a treatment of surface roughness in situations where a perturbative approach would fail. Examples of calculations for surface excitations propagating across the grooves of large-amplitude gratings have been given by Laks et al. (1981) for surface polaritons, Glass and Maradudin (1981) for surface plasmons, and Glass et al. (1981) for Rayleigh waves. For further details of the topic of surface roughness, the reader is referred to appropriate sections of the review articles by Maradudin (1981) and Cottam and Maradudin (1984). 310
Appendix Green functions and linearresponse theory
Here we summarise some basic results concerning Green functions and linear-response functions, and the connection between these quantities. Comprehensive treatments of Green functions are to be found in the numerous text-books on many-body theory; for example, Fetter and Walecka (1971) and Rickayzen (1980). General accounts of the linearresponse method have been given, for example, by Landau and Lifshitz (1969), Barker and Loudon (1972) and Forster (1975), while its specific application to surface problems has been reviewed by Cottam and Maradudin (1984). A.I Basic properties of Green functions For any two quantum-mechanical operators A and B we shall define the Green functions ((A(t); £(£')>>, with time labels t and t\ by «A(t); B(t')}) = i6(t - ?KlA(t), B(f')]e>
(A.I)
Here 6(t — t') is the unit step function corresponding to
*-«•>-{; t
and the operators are in the Heisenberg representation, i.e. A(t) = exp(i^fot)/4 exp(-Uf o r)
(A.3)
where Jf0 is the Hamiltonian describing the system. The angular brackets in (A.I) denote a thermal average, evaluated according to equilibrium statistical mechanics. Finally lA(t)9 B{t')\ = A(t)B(t') - eB(t')A(t)
(A.4)
where e is a parameter which can be chosen as either 1 or — 1. The conventional choice in boson and fermion problems is to take e = 1 and e = — 1, respectively, but this is not essential. 311
Appendix: Green functions and linear-response theory The definition in (A.I) is the same, apart from the overall sign, as that employed by Zubarev (1960) for retarded Green functions. Advanced and causal Green functions can also be introduced but we shall not require them here. One of the simplest properties of < l ( 0 ; * ( * ' ) » to be defined by
-j:
(A.5)
For most cases of practical interest the Green function {{A; !!}}„ can only be evaluated by making approximations, and several different approaches are available. One method is based on writing down the equation of motion of the Green function (Zubarev 1960): (A.6) This generates another Green function, whose equation of motion can itself be written down. After a certain number of stages of this process a decoupling approximation is used to simplify the Green function in the final equation and to yield a closed set of equations which can be solved for the required Green function. An example occurs in §5.3. Another method involves using perturbation theory in a diagrammatic expansion (Feynman diagrams). Details are to be found in the references given earlier. For most of the applications considered in this book we employ yet another approach, namely linear-response theory, which has a more direct physical appeal and enables appropriate Green functions to be deduced using macroscopic arguments. A.2 Linear-response theory A simple illustrative example of the method is given in §1.3.2, and we now present the formal theory. We make use of the concept of the density matrix p (for a review see ter Haar 1961), which may be defined by X™1 I
\
/
p = 2^ \m)pm(jn\
I
/ A *7\
(A.7)
m
where pm is the probability of the system being in the state denoted by |m> (assumed to be orthogonalised and normalised), and the summation is over all states. The density matrix is useful for applications to nonequilibrium (time-dependent) situations. 312
Linear-response theory We consider a system with total Hamiltonian J f given by
j r = JTo + Jfi(0
(A.8)
where Jfo is independent of time and Mf^t) is an external perturbation taking the form jTl(t)=-Bf(t)
(A.9)
Here f(t) denotes an external field which couples linearly with a system variable represented by the operator B, giving the interaction term (A.9). For example, if B represents an atomic displacement then f(t) would be a mechanical force (as in §2.4). We suppose that at time t = — oo the system is in equilibrium, described by the Hamiltonian 3tif0 and the corresponding density matrix Po
= exp(-Jf o /fe B r)/tr[exp(-jr o /* B r)]
(A.10)
where tr denotes the trace of the operators. The perturbation J^x is then increased adiabatically from zero, so that at a later time t the density matrix is equal to P = Po + Pi
(A.11)
The density matrix has the equation of motion (see ter Haar 1961) i dp/dt = tJT, p]
0=1)
(A. 12)
which can be proved from the definition (A.7). If (A.8) and (A.ll) are substituted into (A. 12) and only the terms which are linear in the perturbing field f(t) are retained we find i d P l /dr = [jf0, P l ] - [5, P o ]/(r)
(A.13)
where we have used the property that 3t?0 and p 0 commute. Equation (A.13) can be employed to show that - {exp(iJT0OPi exp(-iJf o r)} = i exp(iJTo0[*, Po] exp(-iJf o ')/(0
(A.14)
On integrating both sides of (A.14) with respect to t and rearranging, the formal solution for px is obtained as
< From the above expression, which is linear in / , we may deduce the response of the system to the perturbation (A.9). This response can be expressed in terms of the change A(t) that it produces in the mean value of any system variable denoted by the operator A. From the definition (A.7) it can be shown that the mean value of A is simply given by tr(p^4). The reader not familiar with this property is invited to work through a 313
Appendix: Green functions and linear-response theory proof in Problem A.4. Hence we have in the present case A(t) = tr(PlA)
(A.16)
On substituting (A.15) into (A.16) and using the cyclic invariance property for products of operators within the trace, we may express A(t) as
where <...> = t r ( P o . . . )
(A.18)
denotes a thermal average with respect to the unperturbed system. When use is made of (A.I), taking the case of a commutator Green function (s= 1), (A.17) becomes
-J:
A{t)=
«A(t);B{t'))>f(t')dt'
(A.19)
J -oo
In terms of frequency Fourier components the above result is simply Io = « A ; B» w F(co)
(A.20)
where {{A; -B>>w is given by (A.5), and f(t) exp(iart) At
(A.21)
0
with a similar definition for I w as the Fourier transform of A(t). Hence by explicitly calculating the response Am for a given externally applied field F(a>) we have a convenient method of deducing from the relationship (A.20) the required Green function <<X; £>>«,. Apart from the frequency the Green function will typically depend on other quantities such as position or wavevector labels, depending on the nature of the problem. For example, in surface problems it is often convenient to assume that the external field takes the form corresponding to a delta-function stimulus f(t)d(r-rf) at position r' within the system. The interaction energy with a position-dependent operator B(T) then becomes •#i = - f B(r)S(r - r') d 3 r/(0 = - B(r')f(t)
(A.22)
From the linear response in another operator A(r) at position r we are able to deduce, by analogy with (A.20), the Green function <> w . Examples for specific systems are given in the text (e.g. in §2.4). A.3 Thefluctuation-dissipationtheorem An important property of Green functions (response functions) is that they give the power spectrum of the excitations of the system. This is 314
The fluctuation-dissipation theorem achieved by application of the fluctuation-dissipation theorem, which relates the mean-square fluctuations in the excitation amplitude to the imaginary part of an appropriately defined Green function. We now outline the quantum-mechanical derivation of this result: for an alternative classical formulation (valid at high temperatures) we refer to Landau and Lifshitz (1969). The Green-function definition in (A. 1) involves the correlation functions CAB(t-t') = (A(t)B(t')) CBA(t-t') = (B(t')A(t)) Their dependence on (t — t') can be proved in the same way as for the Green function. It is easily shown, using the expression (A. 18) for the thermal average and the cyclic invariance property of operators under the trace, that the two correlation functions are related to one another by CAB{t-t') = CBA{t-t' + \p)
(A.24)
where /? denotes l/feBT. If frequency Fourier transforms {AB}^ and {BA}^ are defined for CAB(t — t') and CBA(t — t') respectively, as in (A.21) for F(a>), it is easy to show that (A.24) implies (BA}(O = cxp(M
(A.25)
When both sides of (A.I) are Fourier-transformed to the frequency representation, making use of (A.5) and (A.25), the result is [.-exp(-^)]d,-/ (o-o) +irj
(A.26)
where rj is a positive infinitesimal quantity (rj -* 0). In obtaining the above result it is helpful to use the identity (see Zubarev 1960) 2TTJ_00
(x + irj)
which can be proved by contour integration as in Problem A.5. Equation (A.26) can be separated into its real and imaginary parts by applying the symbolic mathematical relationship 1
)-ind(co-co') (A.28) co — co + irj \co — co' where 0* denotes that the principal value is taken in any integration over co'. Provided (BAy^ is a real quantity (e.g. as in the case when B = A""), we may take the imaginary part of each side of (A.26) to obtain f
Im<04; £ » w = n{\ - exp(-/to>)}<*>0«,
(A.29)
On rearrangement of the terms, this becomes = (l/n){n(co) + 1} Im«X; £ » „
(A.30) 315
Appendix: Green functions and linear-response theory where n(a>) is the Bose-Einstein factor Using (A.25) we also have the result {AB}^ = (l/n)n(a>) lm«A;
B»
w
(A.32)
Equations (A.30) and (A.32) are the standard forms of the flucutationdissipation theorem. At high temperatures (such that /Ja>« 1) they both reduce to the classical limit «
m
= <£M>W = (kBT/nco) I m « > l ; £ » „
(A.33)
By taking the real part of each side of (A.26) we obtain the additional relationship
Re«,;
J-
(©' - c o )
If {1 —exp( —/?«')} can be approximated by /?co' in the classical limit (|j?a/| « 1) we have for the special case of a> = 0 Re
ra=o
=0 J — oo
(A.35)
The integral on the right-hand side of (A.35) represents the total intensity (BA}tot associated with this correlation function, and so we arrive at the simple result that <J5A>tot = kBT R e « 4 ; B)} 0)
=0
(A.36)
Equation (A.36) holds provided the contribution from the region of large \co'\ (such that |/ta/| >> 1) can be neglected for the integral in (A.34). A.4 Problems A.I Derive the Green-function equation of motion in the form quoted in equation (A.6). You should start by differentiating both sides of (A.I) with respect to t, then use the operator equation of motion
and the definition (A.5). A.2 By starting from the definition given in equation (A.7), then differentiating both sides with respect to t and using the time-dependent Schrodinger equation, verify the equation of motion (A. 12) for the density matrix p. A.3 Check that equation (A.14) follows from (A.13). A.4 Prove the stated property (in §A.2) that the mean value of any operator A is equal to tr(pA) where p is the density matrix. You may find it helpful 316
Problems to rewrite tr(pv4) as
I <<>4> i
where the summation is over the complete set of states |i>, and then to employ the definition (A.7) for p. A.5 Derive equation (A.27) by evaluating the integral on the right-hand side and showing that it is equal to 0(t — t') as defined in (A.2). It is convenient to use contour integration, noting that there is a simple pole at x = —iy and the contour of integration may be taken as an infinite semicircle, completed either in the upper or lower half-plane depending on the sign of t-t' (see Fig. A.I). A.6 Verify the algebraic steps leading to equation (A.26).
Fig. A.I The complex x plane and the choice of contours of integration for Problem A.5.
^^\
Contour >v
y / ^^^^^
for / < t'
Contour for / > /'
317
References
Chapter 1 Ashcroft N. W. and Mermin N. D. (1976) Solid State Physics. Holt, Rinehart and Winston: New York Barker A. S. and Loudon R. (1972) Rev. Mod. Phys. 44, 18 Bennett B. I., Maradudin A. A. and Swanson L. R. (1972) Ann. Phys. (NY) 71, 357 Bloch F. (1928) Z. Physik. 52, 555 Brusdeylins G., Doak R. B. and Toennies J. P. (1980) Phys. Rev. Lett. 44, 1417 Brusdeylins G., Doak R. B. and Toennies J. P. (1981) Phys. Rev. Lett. 46, 437 Cottam M. G. (1976) J. Phys. C. 9, 2137 Cottam M. G. and Lockwood D. J. (1986) Light Scattering in Magnetic Solids. Wiley: New York Cottam M. G. and Maradudin, A. A. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 1. North-Holland: Amsterdam Elliott R. J. and Gibson A. F. (1974) An Introduction to Solid State Physics and its Applications. Macmillan: London Felcher G. P. (1985) in Dynamical Phenomena at Surfaces, Interfaces and Super lattices, ed. F. Nizzoli, K. H. Rieder and R. F. Willis, p. 316. Springer: Berlin Hayes W. and Loudon R. (1978) Scattering of Light by Crystals. Wiley: New York Ibach H. and Mills D. L. (1982) Electron Energy Loss Spectroscopy and Surface Vibrations. Academic: New York Joyce B. A. (1973) Surf. Sci. 35, 1 Kittel C. (1986) Introduction to Solid State Physics, 6th edn. Wiley: New York Landau L. D. and Lifshitz E. M. (1969) Statistical Physics. Pergamon: Oxford Laramore G. E. and Switendick A. C. (1973) Phys. Rev. B 7, 3615 Loudon R. (1978) J. Raman Spectrosc. 1, 10 Martin M. R. and Somorjai G. A. (1973) Phys. Rev. B 1, 3607 Mattei G., Pagannone M., Fornari B. and Mattioli L. (1982) Solid State Commun. 44,1495 Mills D. L., Maradudin A. A. and Burstein E. (1970) Ann. Phys. (NY) 56, 504 Monch W. (1973) Adv. Solid State Phys. 13, 241 Nizzoli F. (1986) in Electromagnetic Surface Excitations, ed. R. F. Wallis and G. I. Stegeman, p. 138. Springer: Berlin Nkoma J. S. and Loudon R. (1975) J. Phys. C. 8, 1950 Prutton M. (1983) Surface Physics, 2nd edn. Oxford University Press: Oxford Rayleigh (Lord) (1887) Proc. Lond. Math. Soc. 17, 4 Sandercock J. R. (1970) Optics Commun. 2, 73
318
References Sandercock J. R. (1982) in Light Scattering in Solids III, ed. M. Cardona and G. Giintherodt, p. 173. Springer: Berlin Somorjai G. A. (1975) in Surface Science, Vol. I, p. 173. IAEA: Vienna Szeftel J., Lehwald S. and Ibach H. (1984) J. de Physique 45, C5-109 Tilley D. R. (1980) J. Phys. C. 13, 781 Wherrett B. S. (1986) Group Theory for Atoms, Molecules and Solids. Prentice-Hall: London Whittaker E. T. and Watson G. N. (1963) A Course of Modern Analysis, 4th edn. Cambridge University Press: Cambridge Woodruff D. P. and Delchar T. A. (1986) Modern Techniques of Surface Science. Cambridge University Press: Cambridge Ziman J. M. (1972) Principles of the Theory of Solids, 2nd edn. Cambridge University Press: Cambridge Chapter 2 Albuquerque E. L. (1980) J. Phys. C 13, 2623 Albuquerque E. L., Loudon R. and Tilley D. R. (1980) J. Phys. C 13, 1775 Barber D. J. and Loudon R. (1989) The Macroscopic Properties of Condensed Matter. Cambridge University Press: Cambridge Bortolani V., Nizzoli F., Santoro G., Marvin A. and Sandercock J. R. (1979) Phys. Rev. Lett. 43, 224 Bortolani V., Nizzoli F., Santoro G., Marvin A. and Sandercock J. R. (1979) 43, 224 Bortolani V., Marvin A. M., Nizzoli F. and Santoro G. (1983) J. Phys. C 16, 1757 Bortolani V., Franchini A., Nizzoli F. and Santoro G. (1984) Phys. Rev. Lett. 52, 429 Bortolani V., Franchini A. and Santoro G. (1986) in Electromagnetic Surface Waves, Lecture notes from 1985 Erice Summer School, ed. G. I. Stegeman and R. F. Wallis. Springer: Berlin Brusdeylins G., Doak R. P. and Toennies J. P. (1980) Phys. Rev. Lett. 44, 1417 Brusdeylins G., Rechsteiner R., Skofronick J. G., Toennies J. P., Benedek G. and Miglio L. (1985) Phys. Rev. Lett. 54, 466 Burt M. G. (1973) J. Phys. C 6, 855 Byrne D. and Earnshaw J. C. (1979a) J. Phys. D 12, 1133 Byrne D. and Earnshaw J. C. (1979b) J. Phys. D 12, 1145 Cottam M. G. and Maradudin A. A. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 1. North-Holland: Amsterdam Dervisch A. and Loudon R. (1976) J. Phys. C 9, L669 Doak R. B., Harten U. and Toennies J. P. (1983) Phys. Rev. Lett. 51, 578 Earnshaw J. C. and McGivern R. C. (1987) J. Phys. D 20, 82 Garca-Moliner F. (1977) Ann. Phys. (Paris) 2, 179 Garcia-Moliner F. and Flores F. (1979) Introduction to the Theory of Solid Surfaces. Cambridge University Press: Cambridge Harley R. T. and Fleury P. A. (1979) J. Phys. C 12, L863 Hayes W. and Loudon R. (1978) Scattering of Light by Crystals. Wiley: New York Kueny A., Grimsditch M., Miyano K., Banerjee I., Falco C. M. and Schuller J. K. (1982) Phys. Rev. Lett. 48, 166 Landau L. D. and Lifshitz E. M. (1959) Fluid Mechanics. Pergamon: Oxford Landau L. D. and Lifshitz E. M. (1970) Theory of Elasticity. Pergamon: Oxford Lehwald S., Szeftel J. M., Ibach H., Rahman T. S. and Mills D. L. (1983) Phys. Rev. Lett. 50, 518 Lindsay S. M., Jackson H. E., Harley R. T. and Anderson M. W. (1985) Proceedings of the 17th International Conference on the Physics of Semiconductors, p. 1141. New York: Springer Loudon R. (1978a) J. Phys. C 11, 403 Loudon R. (1978b) J. Phys. C 11, 2623
319
References Loudon R. (1978c) Phys. Rev. Lett. 40, 581 Loudon R. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 589. North-Holland: Amsterdam Loudon R. and Sandercock J. R. (1980) J. Phys. C 13, 2609 Maradudin A. A., Montroll E. W., Weiss G. H. and Ipatova I. P. (1971) Theory of Lattice Dynamics in the Harmonic Approximation. Academic: New York Marvin A. M., Bortolani, V. and Nizzoli F. (1980a) J. Phys. C 13, 299 Marvin A. M., Bortolani, V., Nizzoli F. and Santoro G. (1980b) J. Phys. C 13, 1607 Mazur P. and Maradudin A. A. (1981) Phys. Rev. B 24, 2996 Meeker T. R. and Meitzler A. H. (1964) in Physical Acoustics, ed. W. P. Mason, Vol. 1A, p. 112. Academic: New York Nizzoli F. (1986) in Electromagnetic Surface Waves, Lecture notes from 1985 Erice Summer School, ed. G. I. Stegeman and R. F. Wallis. Springer: Berlin Nye J. F. (1985) Physical Properties of Crystals. Clarendon: Oxford Oliveira F. A., Cottam M. G. and Tilley D. R. (1980) Phys. Status Solidi (b) 107, 737 Oliveros M. C. and Tilley D. R. (1983) Phys. Status Solidi (b) 119, 675 Oshima C , Souda R., Aono M., Otani S. and Ishizawa Y. (1984) Phys. Rev. B 30, 5361 Rowell N. and Stegeman G. I. (1978a) Solid State Commun. 26, 809 Rowell N. and Stegeman G. I. (1978b) Phys. Rev. B 18, 2598 Sandercock J. R. (1972a) Phys. Rev. Lett. 28, 237 Sandercock J. R. (1972b) Phys. Rev. Lett. 29, 1735 Sandercock J. R. (1978) Solid State Commun. 26, 547 Sandercock J. R. (1982) in Light Scattering in Solids III, ed. M. Cardona and G. Guntherodt. Springer: Berlin Scholte J. G. (1947) Mon. Not. R. Astron. Soc: Geophys. Suppl. 5, 120-6 Stegeman G. I. and Nizzoli F. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 195. North-Holland: Amsterdam Subbaswamy K. R. and Maradudin A. A. (1978) Phys. Rev. B 18, 4181 Toennies J. P. (1987) in Surface Phonons, ed. W. Kress. Springer: Berlin Vacher R., Sussner H. and Schmidt M. (1980) Solid State Commun. 34, 279 Wallis R. F. (1957) Phys. Rev. 105, 540 Wallis R. F. (1959) Phys. Rev. 116, 302 Xu M. L., Hall B. M., Tong S. Y., Rocca M., Ibach H., Lehwald S. and Black J. E. (1985) Phys. Rev. Lett. 54, 1171
Chapter 3 Alvarado S. F., Kisker E. and Campagna M. (1986) in Magnetic Properties of LowDimensional Systems, ed. L. M. Falicov and J. L. Moran-Lopez, p. 53. Springer: Heidelberg Anderson P. W. (1963) in Magnetism, Vol. 1, ed. G. T. Rado and H. Suhl, p. 25. Academic: New York Binder K. (1981) Ferroelectrics 35, 99 Blinc R. and Zeks B. (1974) Soft Modes in Ferroelectrics and Antiferroelectrics. NorthHolland: Amsterdam Bloch F. (1930) Z. Phys. 61, 206 Campagna M., Sattler K. and Siegmann H. C. (1974) AIP Conf. Proc. 18, 1388 Celotta R. J., Pierce D. J., Wang G. C , Bader S. D. and Felcher G. P. (1979) Phys. Rev. Lett. 43, 728 Chain's L. J. (1975) in The Helium Liquids, ed. J. G. M. Armitage and I. E. Farquar. Academic: London, p. 473 Cottam M. G. (1976a) J. Phys. C 9, 2121 Cottam M. G. (1976b) J. Phys. C 9, 2137 Cottam M. G. (1978a) J. Phys. C 11, 151
320
References Cottam M. G. (1978b) J. Phys. C 11, 165 Cottam M. G. (1979) J. Phys. C 12, 3541 Cottam M. G. (1983) Solid State Commun. 45, 771 Cottam M. G. and Kontos D. (1980) J. Phys. C 13, 2945 Cottam M. G. and Maradudin A. A. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 1. North Holland: Amsterdam Cottam M. G., Tilley D. R. and 2eks B. (1984) J. Phys. C 17, 1793 Dewames R. E. and Wolfram T. (1969) Phys. Rev. 185, 720 Dos Santos A. R. and Cottam M. G. (1980) Proc. 8th Int. Conf. on Raman Spectroscopy, ed. W. F. Murphy, p. 98. North-Holland: Amsterdam Fillipov B. N. (1967) Sov. Phys. 9, 1048 Harada I., Nagai O. and Nagamiya T. (1977) Phys. Rev. B 16, 4882 Hayakawa K., Namikawa K. and Miyaka S. (1971) J. Phys. Soc. Japan 31, 1408 Hayes W. and Loudon R. (1978) Scattering of Light by Crystals. Wiley: New York Heisenberg W. (1928) Z. Phys. 49, 619 Henderson A. J., Meyer H. and Guggenheim H. J. (1969) Phys. Rev. 185, 128 Henderson A. J., Meyer H. and Guggenheim H. J. (1971) J. Phys. Chem Solids 32, 1047 Herring C. (1966) in Magnetism, Vol. 4, ed. G. T. Rado and H. Suhl. Academic: New York Jarrett H. S. and Waring R. K. (1958) Phys. Rev. I l l , 1223 Kittel C. (1986) Introduction to Solid State Physics, 6th edn. Wiley: New York Levstik A., Tilley D. R. and 2eks B. (1984) J. Phys. C 17, 3793 Levy J. C. S. (1981) Surf. Sci. Rep. 1, 39 Lines M. E. and Glass A. M. (1979) Principles and Applications of Ferroelectrics and Related Materials. Oxford University Press, Oxford Mattis D. C. (1965) The Theory of Magnetism. Harper and Row: New York Mills D. L. (1968) Phys. Rev. Lett. 20, 18 Mills D. L. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 379. North-Holland: Amsterdam Mills D. L. and Beal-Monod M.-T. (1974a) Phys. Rev. A 10, 343 Mills D. L. and Beal-Monod M.-T. (1974b) Phys. Rev. A 10, 2473 Mills D. L. and Maradudin A. A. (1967) J. Phys. Chem. Solids 28, 1855 Mills D. L. and Saslow W. M. (1968) Phys. Rev. 171, 488 Moul R. C. and Cottam M. G. (1979) J. Phys. C 12, 5191 Moul R. C. and Cottam M. G. (1983) J. Phys. C 16, 1307 Palmberg P. W., Dewames R. E. and Viedevoe L. A. (1968) Phys. Rev. Lett. 21, 682 Panofsky W. K. H. and Phillips M. (1962) Classical Electricity and Magnetism, 2nd edn. Addison-Wesley: Reading Phillips T. G. and Rosenberg H. M. (1966) Rep. Prog. Phys. 29, 285 Pierce D. T. (1986) in Magnetic Properties of Low-Dimensional Systems, ed. L. M. Falicov and J. L. Moran-Lopez, p. 58. Springer: Heidelberg Puszkarskii H. (1970) Ada Phys. Polon. A 38, 217 Puszkarskii H. (1972) Phys. Status Solidi (b) 50, 87 Puszkarskii H. (1979) Prog. Surf. Sci. 9, 191 Rado G. T. and Weertman J. R. (1959) J. Phys. Chem. Solids 11, 315 Seavey M. H. and Tannenwald P. E. C. (1958) Phys. Rev. Lett. 1, 168 Schiff L. I. (1955) Quantum Mechanics. McGraw-Hill: New York Wagner D. (1972) Introduction to the Theory of Magnetism. Pergamon: Oxford Wallis R. F., Maradudin A. A., Ipatova I. P. and Klochikhim A. A. (1967) Solid State Commun. 5, 89 Weiss P. (1907) J. Phys. 6, 661 Wolfram T. and Dewames R. E. (1969) Phys. Rev. 185, 762 Wolfram T. and Dewames R. E. (1972) Prof. Surf. Sci. 2, 233 Yu J. T., Turk R. A. and Wigen P. E. (1975) Phys. Rev. B 11, 420
321
References Chapter 4 Brundle L. K. and Freedman N. J. (1968) Electron. Lett. 4, 132 Camley R. E. (1980) Phys. Rev. Lett. 45, 283 Camley R. E. and Maradudin A. A. (1982) Solid State Commun. 41, 585 Camley R. E. and Mills D. L. (1978) Phys. Rev. B 18, 4821 Camley R. E., Rahman T. S. and Mills D. L. (1981) Phys. Rev. B 23, 1226 Cohen M. H. and Keffer F. (1955) Phys. Rev. 99, 1128 Cottam M. G. (1979) J. Phys. C 12, 1709 Cottam M. G. (1983) J. Phys. C 16, 1573 Cottam M. G. and Lockwood D. J. (1986) Light Scattering in Magnetic Solids. Wiley: New York Cottam M. G. and Maradudin A. A. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 1. North-Holland: Amsterdam Damon R. W. and Eshbach J. R. (1961) J. Phys. Chem. Solids 19, 308 Grimsditch M., Malozemoff A. P. and Brunsch A. (1979) Phys. Rev. Lett. 43, 711 Grunberg P. (1980) J. Appl. Phys. 51, 4338 Grunberg P. (1981) J. Appl. Phys. 52, 6824 Grunberg P. (1985) Prog. Surf. Sci. 18, 1 Grunberg P. and Metawe F. (1977) Phys. Rev. Lett. 39, 1561 Grunberg P., Cottam M. G., Vach W, Mayr C. M. and Camley R. E. (1982) J. Appl. Phys. 53, 2078 Ishak W. S. and Chang K.-W. (1985) Hewlett-Packard Journal 36, no. 2, 10 Kabos P., Wilber W. D., Patton C. E. and Grunberg P. (1984) Phys. Rev. B 29, 6396 Keffer F. (1966) Handbuch der Physik 18/2, 1 Loudon R. and Pincus P. (1963) Phys. Rev. 132, 673 Luthi B., Mills D. L. and Camley R. E. (1983) Phys. Rev. B 28, 1475 Mercereau J. and Feynman R. P. (1956) Phys. Rev. 104, 63 Mills D. L. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 379. North-Holland: Amsterdam Sandercock J. R. (1982) in Light Scattering in Solids III, ed. M. Cardona and G. Guntherodt, p. 173. Springer: New York Sandercock J. R. and Wettling W. (1979) J. Appl. Phys. 50, 7784 Sarmento E. F. and Tilley D. R. (1982) in Electromagnetic Surface Modes, ed. A. D. Boardman. Wiley: New York Stamps R. L. and Camley R. E. (1984) J. Appl. Phys. 56, 3497 Tarasenko V. V. and Kharitonov V. D. (1971) Zh. Eksp. Teor. Fiz. 60, 2321 [Sov. Phys.—JETP 33, 1246 (1971)] Tilley D. R. (1986) in Electromagnetic Surface Excitations, ed. R. F. Wallis and G. I. Stegeman, p. 30. Springer: Heidelberg Vernon S. P., Lindsay S. M. and Stearns M. B. (1984) Phys. Rev. B 29, 4439 Walker L. R. (1957) Phys. Rev. 105, 390 Wolfram T. and Dewames R. E. (1970) Phys. Rev. B 1, 4358 Wolfram T. and Dewames R. E. (1972) Prog. Surf. Sci. 2, 233 Chapter 5 Abrikosov A. A., Gorkov L. P. and Dzyaloshinski I. E. (1963) Methods of Quantum Field Theory in Statistical Physics. Prentice-Hall: New Jersey Agranovich V. M. and Ginzburg V. L. (1984) Crystal Optics with Spatial Dispersion, and Excitons. Springer: Berlin Ashcroft N. W. and Mermin N. D. (1976) Solid State Phys. Holt, Rinehart and Winston: New York Boardman A. D. (1982) in Electromagnetic Surface Modes, ed. A. D. Boardman. Wiley: Chichester
322
References Born M. and Huang K. (1985) Dynamical Theory of Crystal Lattices. Clarendon: Oxford Chiaradia P., Cricenti A., Selci S. and Chiarotti G. (1984) Phys. Rev. Lett. 52, 1145 Cole M. W. (1974) Rev. Mod. Phys. 46, 451 Dasgupta B. B. and Beck D. E. (1982) in Electromagnetic Surface Modes, ed. A. D. Boardman. Wiley: Chichester Del Sole R. (1986) in Electromagnetic Surface Excitations, ed. R. F. Wallis and G. I. Stegeman. Springer: Berlin Eastman D. E. and Grobman W. D. (1972) Phys. Rev. Lett. 28, 1378 Eastman D. E. and Freeouf J. L. (1974) Phys. Rev. Lett. 33, 1601 Eastman D. E. and Nathan M. I. (1975) Physics Today 28 (April), 44 Ehrenreich H., Seitz F. and Turnbull D. (1980) Solid State Physics, Vol. 35. Academic: New York Elliott R. J. and Gibson A. F. (1978) An Introduction to Solid State Physics. Macmillan: London Fetter A. L. (1973) Ann. Phys. 81, 367 Fetter A. L. (1974) Ann. Phys. 88, 1 Fetter A. L. and Walecka J. D. (1971) Quantum Theory of Many-Particle Systems. McGraw-Hill: New York Flores F. and Garcia-Moliner F. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon. North-Holland: Amsterdam Forstmann F. (1970) Z. Physik 235, 69 Goodwin E. T. (1939) Proc. Camb. Phil. Soc. 35, 205, 221, 232 Heine V. (1980) Solid State Physics, Vol. 35, ed. H. Ehrenreich, F. Seitz and D. Turnbull, p. 1. Academic: New York Karsono A. D. and Tilley D. R. (1977) J. Phys. C 10, 2123 Kelly M. J. (1980) Solid State Physics, Vol. 35, ed. H. Ehrenreich, F. Seitz and D. Turnbull, p. 296. Academic: New York Kim O. K. and Spitzer W. G. (1979) J. Appl. Phys. 50, 4362 Kittel C. (1986) Introduction to Solid State Physics, 6th edn. Wiley: New York Landau L. D. and Lifshitz E. M. (1971) Electrodynamics of Continuous Media. Pergamon: Oxford Lindhard J. (1954) Kgl. Danske Mat.-fys. Medd. 28, 8 Nozieres P. and Pines D. (1958) Nuovo Cimento [x] 9, 470 Palik E. D., Kaplan R., Gammon R. W., Kaplan H., Wallis R. F. and Quinn J. J. (1976) Phys. Rev. B 13, 2497 Park R. L. (1975) Physics Today 28 (April), 52 Pendry J. B. and Gurman S. J. (1975) Surf. Sci. 49, 87 Plummer E. W., Gadzuk J. W. and Penn D. R. (1975) Physics Today 28 (April), 63 Rickayzen G. D. (1980) Green Functions and Condensed Matter. Academic: London Schrieffer J. R. and Soven P. (1975) Physics Today 28 (April), 24 Shockley W. (1939) Phys. Rev. 56, 317 Stern F. (1967) Phys. Rev. Lett. 18, 546 Tamm I. (1932) Physik. Z. Sowjetunion 1, 733 Tilley D. R. and Tilley J. (1986) Superfluidity and Superconductivity, 2nd edn. Adam Hilger: Bristol Chapter 6 Abeles F. (1986) Electromagnetic Surface Excitations, ed. R. F. Wallis and G. I. Stegeman, p. 8. Springer: Berlin Agranovich V. M. and Mills D. L. (1982) Surface Polaritons. North-Holland: Amsterdam Auld B. A. (1960) J. Appl. Phys. 31, 1642 Barker A. S. and Loudon R. (1972) Rev. Mod. Phys. 44, 18 Batke E. and Heitmann D. (1984) Infrared Physics 24, 189
323
References Boardman A. D. (ed.) (1982) Electromagnetic Surface Modes. Wiley: Chichester Boardman A. D. and Egan D. (1985) IEEE J. Quant. Electron. QE 21, 1701 Borstel G., Falge H. J. and Otto A. (1974) Springer Tracts in Modern Physics 74, 107 Borstel G. and Falge H. J. (1977) Phys. Status Solidi (b) 83, 11 Borstel G. and Falge H. J. (1978) Appi Phys. 16, 211 Bose M. S., Foo E. N. and Zuniga M. A. (1975) Phys. Rev. B 12, 3885 Brenig W., Zeyher R. and Birman J. L. (1972) Phys. Rev. B 6, 4617 Brodin M. S., Bardura V. M. and Matsko M. G. (1984) Phys. Status Solidi (b) 125, 613 Bryksin V. V., Mirlin D. N. and Firsov Yu. A. (1974) Usp. Fiz. Nauk 113, 29 (Sov. Phys. Usp. 17, 305) Burstein E. and De Martini F. (1974) Polaritons. Pergamon: New York Camley R. E. and Mills D. L. (1982) Phys. Rev. B 26, 1280 Chen Y. J., Burstein E. and Mills D. L. (1975) Phys. Rev. Lett. 34, 1516 Cottam M. G. and Maradudin A. A. (1984) Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 1. North-Holland: Amsterdam Cottam M. G., Tilley D. R. and 2eks B. (1984) J. Phys. C 17, 1793 Craig A. E., Olson G. A. and Sarid D. (1983) Opt. Lett. 8, 380 Dawson P., Walmsley D. G., Quinn H. A. and Ferguson A. J. L. (1984) Phys. Rev. B 30,3164 De Martini F., Colocci M., Kohn S. E. and Shen Y. R. (1977) Phys. Rev. Lett. 38, 1223 Evans D. J., Ushioda S. and McMullen J. D. (1973) Phys. Rev. Lett. 31, 372 Fukui M., So V. C. Y. and Normandin R. (1979) Phys. Status Solidi (b), 91, K61 Fukui M., So V. C. Y. and Stegeman G. I. (1980) Phys. Rev. B 22, 1010 Fukui M. and Tada O. (1982) J. Phys. Soc. Japan 51, 172 Fukui M., Dohi H., Matsuura J. and Tada O. (1984) J. Phys. C 17, 1783 Hartstein A., Burstein E., Maradudin A. A., Brewster R. and Wallis R. F. (1973) J. Phys. C 6, 1266 Hayes W. and Loudon R. (1978) Scattering by Light of Crystals. Wiley: New York Kaplan A. E. (1977) Sov. Phys. JETP 45, 896 Karsono A. D. and Tilley D. R. (1978) J. Phys. C 11, 3487 Kim O. K. and Spitzer W. G. (1979) J. Appl. Phys. 50, 4362 Kittel C. (1986) Introduction to Solid State Physics, 6th edn. Wiley: New York Kretschmann E. and Raether H. (1968) Z. Naturf. A 23, 615 Lagois J. and Fischer B. (1976) Phys. Rev. Lett. 36, 680 Lagois J. and Fischer B. (1978) Phys. Rev. B 17, 3814 Lagois J. and Fischer B. (1982) in Surface Polaritons, ed. V. M. Agranovich and D. L. Mills. North-Holland: Amsterdam Landau L. D. and Lifshitz E. M. (1971) Electrodynamics of Continuous Media. Pergamon: Oxford Lima N. P. and Oliveira F. A. (1986) J. Phys. C 19, 5381 Manohar C. and Venkataraman G. (1972) Phys. Rev. B 5, 1993 Maradudin A. A. (1982) in Surface Polaritons, ed. V. M. Agranovich and D. L. Mills, p. 405. North-Holland: Amsterdam Maradudin A. A. (1986) in Electromagnetic Surface Excitations, ed. R. F. Wallis and G. I. Stegeman, p. 57. Springer: Berlin Maradudin A. A. and Mills D L. (1973) Phys. Rev. B 7, 2787 Marchand M. and Caille A. (1980) Solid State Commun. 34, 827 Marcuse D. (1982) Light Transmission Optics, 2nd. edn. van Nostrand: New York Marschall N. and Fischer B. (1972) Phys. Rev. Lett. 28, 811 Matsuura J. Fukui M. and Tada O. (1983) Solid State Commun. 45, 157 Mills D. L., Chen Y. J. and Burstein E. (1976) Phys. Rev. B 13, 4419 Mirlin D. N. (1982) in Surface Polaritons, ed. V. M. Agranovich and D. L. Mills. North-Holland: Amsterdam Nakayama M. (1974) J. Phys. Soc. Japan 36, 393
324
References Nkoma J. S., Loudon R. and Tilley D. R. (1974) J. Phys. C 7, 3547 Nkoma J. S. (1975) J. Phys. C 8, 3919 Nkoma J. S. and Loudon R. (1975) J. Phys. C 8, 1950 Nye J. F. (1985) Physical Properties of Crystals. Clarendon: Oxford Oliveira F. A., Khater A. F., Sarmento E. F. and Tilley, D. R. (1979) J. Phys. C 12, 4021 Otto A. (1974) Festkorperprobleme XIV, 1 Otto A. (1976) Optical Properties of Solids: New Developments, ed. B. O. Seraphim, p. 678. North-Holland: Amsterdam Palik E. D., Kaplan R., Gammon R. W., Kaplan H., Wallis R. F. and Quinn J. J. (1976) Phys. Rev. B 13, 2497 Pines D. and Nozieres P. (1966) The Theory of Quantum Liquids. Benjamin: New York Prieur J.-Y. and Ushioda S. (1975) Phys. Rev. Lett. 34, 1012 Raether H. (1982) in Surface Polaritons, ed. V. M. Agranovich and D. L. Mills, p. 331. North-Holland: Amsterdam Rashba E. I. and Sturge M. D. (1982) Excitons. North-Holland: Amsterdam Remer L., Luthi B., Sauer H., Geick R. and Camley R. E. (1986) Phys. Rev. Lett. 56,2752 Rimbey P. R. and Mahan G. D. (1974) Solid State Commun. 15, 35 Sarid D. (1981) Phys. Rev. Lett. 47, 1927 Sarmento E. F. and Tilley D. R. (1976) J. Phys. C 9, 2943 Sarmento E. F. and Tilley D. R. (1977) J. Phys. C 10, 795 Sarmento E. F. and Tilley D. R. (1982) in Electromagnetic Surface Modes, ed. A. D. Boardman, p. 633. Wiley: New York Smith R. A. (1978) Semiconductors. Cambridge University Press: Cambridge So V. C. Y., Fukui M., Thomas P. J. and Stegeman G. I. (1981) J. Phys. C 14, 4505 Stegeman G. I. and Wallis R. F. (1986) Electromagnetic Surface Excitations. Springer: Berlin Stegeman G. I., Seaton C. T., Hetherington W. M. Ill, Boardman A. D. and Egan P. (1986) in Electromagnetic Surface Excitations, ed. R. F. Wallis and G. I. Stegeman, p. 261. Springer: Berlin Stern F. (1967) Phys. Rev. Lett. 18, 546 Subbaswamy K. R. and Mills D. L. (1978) Solid State Commun. 27, 1085 Tilley D. R. (1980) J. Phys. C 13, 781 Tilley D. R. (1986) in Electromagnetic Surface Excitations, ed. R. F. Wallis and G. I. Stegeman, p. 30. Springer: Berlin Tokura Y., Koda T., Hirabayashi I. and Nakada S. (1981) J. Phys. Soc. Japan 50, 145 Ushioda S. and Loudon R. (1982) in Surface Polaritons, ed. V. M. Agranovich and D. L. Mills. North-Holland: Amsterdam Valdez J. B. and Ushioda S. (1977) Phys. Rev. Lett. 38, 1088 Weisbuch C. and Ulbrich R. G. (1982) in Topics in Applied Physics 51: Light Scattering in Solids III, ed. M. Cardona and G. Guntherodt, p. 207. Springer: Berlin Wendler L. (1986) Phys. Status Solidi (b) 135, 759 Wendler L. and Haupt R. (1986a) Phys. Status Solidi (b) 137, 269 Wendler L. and Haupt R. (1986b) J. Phys. C 19, 1871 Wood R. W. (1935) Phys. Rev. 48, 928 Chapter 7 Agranovich V. M. and Kravtsov V. E. (1985) Solid State Commun. 55, 85 Albuquerque E. L., Fulco P. and Tilley D. R. (1986a) Rev. Bras. Fis 16, 315 Albuquerque E. L., Fulco P. and Tilley D. R. (1988) Phys. Status Solidi (b) 146, 449 Albuquerque E. L.. Fulco P., Sarmento E. F. and Tilley D. R. (1986b) Solid State Commun. 58, 41 Almeida N. S. and Mills D. L. (1988) Phys. Rev. B 37, 3400 Babiker M. and Tilley D. R. (1984) J. Phys. C 17, L829
325
References Babiker M., Tilley D. R., Albuquerque E. L. and Goncalves da Silva C. E. T. (1985a) J. Phys. C 18, 1269 Babiker M., Tilley D. R. and Albuquerque E. L. (1985b) J. Phys. C 18, 285 Babiker M., Constantinou N. C. and Cottam M. G. (1986a) Solid State Commun. 57, 877 Babiker M., Constantinou N. C. and Cottam M. G. (1986b) J. Phys. C 19, 5849 Babiker M., Constantinou N. C. and Cottam M. G. (1986c) Solid State Commun. 59, 751 Babiker M., Constantinou N. C. and Cottam M. G. (1987a) J. Phys. C 20, 4581 Babiker M., Constantinou N. C. and Cottam M. G. (1987b) J. Phys. C 20, 4597 Barnas J. (1987) Solid State Commun. 61, 405 Barnas J. (1988) J. Phys. C 21, 1021 Bastard G. (1981) Phys. Rev. B 24, 5693 Birch J. R. and Parker T. J. (1979) in Infrared and Millimetre Waves, ed. K. J. Button, ch. 3. Academic: New York Bloss W. L. and Brody E. M. (1982) Solid State Commun. 43, 523 Bulgakov A. A. (1985) Solid State Commun. 55, 869 Camley R. E. (1987) Phys. Rev. B 35, 3608 Camley R. E. and Cottam M. G. (1987) Phys. Rev. B 35, 189 Camley R. E. and Mills D. L. (1984) Phys. Rev. B 29, 1695 Camley R. E. and Tilley D. R. (1988) Phys. Rev. B 37, 3413 Camley R. E., Djafari-Rouhani B., Dobrzynski L. and Maradudin A. A. (1983a) Phys. Rev. B 27, 7318 Camley R. E., Rahman T. S. and Mills D. L. (1983b) Phys. Rev. B 27, 261 Chang L. L. and Ploog K. (eds.) (1985) Molecular Beam Epitaxy and Heterostructures. NATO/Martinas Nijhoof: Dordrecht Colvard C , Merlin R., Klein M. V. and Gossard A. C. (1980) Phys. Rev. Lett. 45, 298 Colvard C , Gant T. A., Klein M. V., Merlin R., Fischer R., Morkoc H. and Gossard A. C. (1985) Phys. Rev. B 31, 2080 Constantinou N. C. and Cottam M. G. (1986) J. Phys. C 19, 739 Cottam M. G. and Lockwood D. J. (1986) Light Scattering in Magnetic Solids. Wiley: New York Das Sarma S. and Quinn J. J. (1982) Phys. Rev. B 25, 7603 de Gennes P. G. (1974) The Physics of Liquid Crystals. Clarendon: Oxford Djafari-Rouhani B. and Dobrzynski L. (1987) Solid State Commun. 62, 609 Djafari-Rouhani B., Dobrzynski L., Hardouin-Duparc O., Camley R. E. and Maradudin A. A. (1983) Phys. Rev. B 28, 1711 Djafari-Rouhani B., Sapriel J. and Bonnouvrier F. (1985) Superlattices and Microstructures 1,29 Dobrzynski L., Djafari-Rouhani B. and Puszkarski H. (1986) Phys. Rev. B 33, 3251 Dresselhaus M.S. (1986) in Magnetic Properties of Low-Dimensional Systems, ed. L. M. Falicov and J. L. Moran-Lopez, p. 172. Springer: Heidelberg Duffield T., Bhat R., Koza M., De Rosa F., Hwang D. M., Grabbe P. and Allen S. J. (1986) Phys. Rev. Lett. 56, 2724 Eliasson G., Hawrylak P. and Quinn J. J. (1987) Phys. Rev. B 35, 5569 Fasol G., Hughes H. P. and Ploog K. (1985) Proc. Int. Conf. on Electronic Properties of Two-Dimensional Systems (Tokyo), p. 742 Feldman D. W., Parker J. H., Choyke W. J. and Patrick L. (1968) Phys. Rev. 15, 698 Fetter A. L. (1974) Ann. Phys. (NY) 88, 1 Fishman F., Schwabl F. and Schwenk D. (1987) Phys. Lett. A 121, 192 Giuliani G. F. and Quinn J. J. (1983) Phys. Rev. Lett. 51, 919 Giuliani G. F., Qin G. and Quinn J. J. (1984) Surf. Sci. 142, 433 Grimsditch M., Khan M. R., Kueny A. and Schuller I. K. (1983) Phys. Rev. Lett. 51, 498 Grunberg P. and Mika K. (1983) Phys. Rev. B 27, 2955 Hartstein A., Burstein E., Brion J. J. and Wallis R. F. (1973) Solid State Commun. 12,1083
326
References Haupt R. and Wendler L. (1987) Solid State Commun. 612, 341 Hawrylak P., Wu J. W. and Quinn J. J. (1985) Phys. Rev. B 32, 5169 Hinchey L. L. and Mills D. L. (1986a) Phys. Rev. B 33, 3329 Hinchey L. L. and Mills D. L. (1986b) Phys. Rev. B 34, 1689 Jain J. K. and Allen P. B. (1985a) Phys. Rev. Lett. 54, 947 Jain J. K. and Allen P. B. (1985b) Phys. Rev. Lett. 54, 2437 Jain J. K. and Allen P. B. (1985c) Phys. Rev. B 32, 997 Joyce B. A. (1985) Rep. Prog. Phys. 48, 1637 Jusserand B., Paquet D., Regreny A. and Kervarec J. (1983) Solid State Commun. 48, 499 Jusserand B., Paquet D., Kervarec J. and Regreny A. (1984a) J. de Physique 45, C5-154 Jusserand B., Paquet D. and Regreny A. (1984b) Phys. Rev. B 30, 6245 Jusserand B., Paquet D. and Regreny A. (1985) Superlattices and Microstructures 1, 61 Katayama S. and Ando T. (1985) J. Phys. Soc. Japan 54, 1615 King-Smith R. D. and Inkson J. C. (1986) Phys. Rev. B 33, 5489 King-Smith R. D. and Inkson J. C. (1987) Phys. Rev. B 36, 4796 Klein M. V. (1986) IEEE Journal Quantum Electronics QE-22, 1760 Kueny A., Khan M. R., Schuller I. K. and Grimsditch M. (1984) Phys. Rev. B 29, 2879 Landau L. D. and Lifshitz E. M. (1969) Statistical Physics. Pergamon: Oxford Lipson S. G. and Lipson H. (1969) Optical Physics. Cambridge University Press: Cambridge Liu W.-M., Eliasson G. and Quinn J. J. (1985) Solid State Commun. 55, 533 Maslin K. A., Parker T. J., Raj N., Tilley D. R., Dobson P. J., Hilton D. and Foxon C. T. B. (1986) Solid State Commun. 60, 461 Olego D., Pinczuk A., Gossard A. C. and Wiegmann W. (1982) Phys. Rev. B 25, 7867 Parker E. H. C. (ed.) (1985) The Technology and Physics of Molecular Beam Epitaxy. Plenum: New York Pinczuk A. (1984) J. de Physique 45, C5-477 Qin G., Giuliani G. F. and Quinn J. J. (1983) Phys. Rev. B 28, 6144 Raj N. and Tilley D. R. (1985) Solid State Commun. 55, 373 Raj N. and Tilley D. R. (1987) Phys. Rev. B. 36, 7003 Raj N., Camley R. E. and Tilley D. R. (1987) J. Phys. C 20, 5203 Rytov S. M. (1956) Akust. Zh. 2, 71 \_Sov. Phys.—Acoust. 2, 68 (1956)] Sooryakumar R., Pinczuk A., Gossard A. C. and Wiegeman W. (1985) Phys. Rev. B 31,2578 Taborelli M., Allenspach R., Boffa G. and Landolt M. (1986) Phys. Rev. Lett. 56, 2869 Tsellis A. and Quinn J. J. (1984) Phys. Rev. B 29, 3318 Wallis R. F. (1957) Phys. Rev. 105, 540 Wasserman A. L. and Lee Y. I. (1985) Solid State Commun. 54, 855 Yariv A. and Yeh P. (1977) J. Opt. Soc. Amer. 67, 438 Yariv A. and Yeh P. (1984) Optical Waves in Crystals. Wiley: New York Yeh P., Yariv A. and Hong C.-S. (1977) J. Opt. Soc. Amer. 67, 423 Yip S. K. and Chang Y. C. (1984) Phys. Rev. B 30, 7037 Zucker J. E., Pinczuk A., Chemla D. J., Gossard A. and Wiegmann W. (1984) Phys. Rev. Lett. 53, 1280 Zucker J. E., Pinczuk A., Chemla D. J., Gossard A. and Wiegmann W. (1985) Proc. 17th Int. Conf. on the Physics of Semiconductors, ed. J. D. Chadi and W. A. Harrison, p. 563. Springer: New York Chapter 8 Boardman A. D., Aers G. C. and Teshima R. (1981) Phys. Rev. B 24, 5703 Burstein E., Lundquist S. and Mills D. L. (1982) in Surface Enhanced Raman Scattering, ed. R. K. Chang and T. E. Furtak, p. 67. Plenum: New York Camley R. E. and Fulde P. (1981) Phys. Rev. B 23, 2614 Cottam M. G. and Maradudin A. A. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 1. North-Holland: Amsterdam
327
References Dobrzynski L. and Maradudin A. A. (1972) Phys. Rev. B 6, 3810 Economou E. N. and Ngai K. L. (1974) Adv. Chem. Phys. 27, 265 Englman R. and Ruppin R. (1968a) J. Phys. C 1, 614 Englman R. and Ruppin R. (1968b) J. Phys. C 1, 1515 Fletcher P. C. and Kittel C. (1960) Phys. Rev. 120, 2004 Glass N. E., Loudon R. and Maradudin A. A. (1981) Phys. Rev. B 24, 6843 Glass N. E. and Maradudin A. A. (1981) Phys. Rev. B 24, 595 Hudson G. E. (1943) Phys. Rev. 63, 46 Joseph R. I. and Schlomann E. (1961) J. Appl. Phys. 32, 1001 Kittel C. (1958) Phys. Rev. 110, 836 Kliewer K. L. and Fuchs R. (1974) Adv. Chem. Phys. 27, 355 Kontos D. and Cottam M. G. (1984) J. de Physique 45, C5-329 Kontos D. and Cottam M. G. (1986) J. Phys. C 19, 1189 Kretschmann E., Ferrell T. L. and Ashley J. C. (1979) Phys. Rev. Lett. 42, 1312 Lagasse P. E. (1972) Electron. Lett. 8, 372 Laks B., Mills D. L. and Maradudin A. A. (1981) Phys. Rev. B 23, 4965 Lingner C. and Luthi B. (1981) Phys. Rev. B 23, 256 Maradudin A. A. (1981) in Festkorperprobleme (Advances in Solid State Physics); Vol. 21, ed. J. Treusch, p. 25. Vieweg: Braunschweig Maradudin A. A., Moss S. L. and Cunningham S. L. (1977) Phys. Rev. B 15, 4490 Maradudin A. A., Wallis R. F., Mills D. L. and Ballard R. L. (1972) Phys. Rev. B. 6, 1106 Mazur P. and Mills D. L. (1984) Phys. Rev. B 29, 5081 Mercerau J. and Feynman R. P. (1956) Phys. Rev. 104, 63 Moss S. L., Maradudin A. A. and Cunningham S. L. (1973) Phys. Rev. B 8, 2999 Palmer R. E. and Schnatterly S. E. (1971) Phys. Rev. B 4, 2329 Rischbieter F. (1965) Acustica 16, 75 Rischbieter F. (1967) Acustica 18, 109 Ruppin R. and Englman R. (1968) J. Phys. C 1, 631 Scott R. Q. and Mills D. L. (1977) Phys. Rev. B 15, 3545 Sharon T. M. and Maradudin A. A. (1973) Solid State Commun. 13, 187 Stegeman G. I. and Nizzoli F. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 195. North-Holland: Amsterdam Tarasenko V. V. and Kharitonov V. D. (1974) Sov. Phys.—Solid State 16, 1031 Urazokov E. I. and Falkovskii L. A. (1972) Zh. Eksp. i Teor. Fiz. 63, 2297 [Sov. Phys.—JETP 36, 1214 (1973)] Ushioda S., Aziza A., Valdez J. B. and Mattei G. (1979) Phys. Rev. B 19, 4012 Viktorev I. A. (1963) Akust. Zh. 9, 296 [Sov. Phys.—Acoustics 9, 242 (1964)] Walker L. R. (1957) Phys. Rev. 105, 390 Wallace D. C. (1970) in Solid State Physics, Vol. 25, ed. H. Ehrenreich, F. Seitz and D. Turnbull, p. 301. Academic: New York Wolfram T. and Dewames R. E. (1972) Prog. Surf. Sci. 2, 233 Appendix Barker A. S. and Loudon R. (1972) Rev. Mod. Phys. 44, 18 Cottam M. G. and Maradudin A. A. (1984) in Surface Excitations, ed. V. M. Agranovich and R. Loudon, p. 1. North-Holland: Amsterdam Fetter A. L. and Walecka J. D. (1971) Quantum Theory of Many-Particle Systems. McGraw-Hill: New York Forster D. (1975) Hydrodynamic Fluctuations, Broken Symmetry and Correlation Functions. Benjamin: New York Landau L. D. and Lifshitz E. M. (1969) Statistical Physics, Pergamon: Oxford Rickayzen G. (1980) Green's Functions and Condensed Matter. Academic: London ter Haar D. (1961) Rep. Prog. Phys. 24, 304 Zubarev D. N. (1960) Usp. Fiz. Nauk 71, 71 [Sov. Phys.—Usp. 3, 320 (I960)]
328
Index
ABCs, see additional boundary conditions acoustic phonons; see Phonons acousto-optic tensor 58ff, 66ff additional boundary conditions electron-gas 172, 175f excitonic 190ff magnetic 139 AES, see Auger electron spectroscopy Ag 69, 70, 171,223,228, 234 Al 66 antiferromagnets bulk magnetostatic modes 145 bulk magnons llOff, 115 bulk polaritons 236 Hamiltonian 109 spin-flop transition 116 surface magnetostatic modes 142ff surface magnons 113ff, 116 surface polaritons 240 antiferromagnetic resonance 111, 145 anti-Stokes scattering definition 24 magnons 152, 148, 150, 289 via exciton-polaritons 193 see also Brillouin scattering; Raman scattering attenuated total reflection (ATR): angle scan 221 basic principles 218ff frequency scan 221 Otto configuration 219ff, 223 Raether-Kretschmann configuration 219ff superlattices 270 surface exciton-polaritons 225f theory 222ff attenuation length 9, 12 Au 225 Auger effect 23, 163 Auger electron spectroscopy 12, 22f
band offset 272 basis 2 Bloch's theorem 4,9,34,36,93,159,219, 251, 264, 273, 278 boundary conditions acoustic 36, 43, 250 cyclic 6 electromagnetic 196, 203, 210, 264, 278 electron 159, 272 electrostatic 201 liquid surface 77 magnons 98 see also additional boundary conditions Bravais lattice 7 Brillouin scattering acoustic surface modes 56ff double layer ferromagnets 152 ferromagnetic films 150ff magnetic superlattices 289 principles 23ff resonant 188ff, 192ff Brillouin zone 4, 8 bulk continuum 47, 94, 146f causality 18 central peak 64 confined modes 260, 262, 274, 298 constant-amplitude approximation 295 correlation functions 312f Curie temperature 86 cyclic boundary conditions, see boundary conditions, cyclic cyclotron frequency 176, 274 cylindrical surfaces 3O7ff Damon-Eshbach modes, see magnetostatic modes, ferromagnetic dangling bonds 14 dead-layer model 191 decay length; see attenuation length
329
Index decoupling approximation 168, 312 see also random phase approximation demagnetising factor 129 density matrix 312 DFTS, see dispersive Fourier transform spectroscopy diatomic lattice, see lattice dynamics dielectric function definition 163f electron gas 169ff excitonic 189 intensity-dependent 215 ionic crystals 177ff longitudinal 165 RPA 170 semiconductors 179 superlattice 267 transverse 165 differential cross section definition 27 formula 32, 61 light scattering by magnons 102 dipole-dipole interaction 87, 128ff dipole-exchange modes, ferromagnets 137ff, 142 region 127 dispersive Fourier transform spectroscopy 267f, 276 displacement derivatives 59 edge modes 3O5ff EELS, see electron energy loss spectroscopy elasticity tensor 41 elasticity theory linear 40ff nonlinear 304 elastic waves bulk 10,42ff single interface 10, 45ff, 304 superlattices 249ff two-interface 47ff see also phonons, acoustic electromagnetic region 127 electron diffraction low-energy 12, 22, 118 reflection high-energy 22 spin-polarised 118 electron energy loss spectroscopy, principles 28f electron gas collective properties 165ff compressibility 174 dielectric function 169ff hydrodynamic description 174ff, 177 superlattices two-dimensional 176f, 203f, 276f electron states collective, see plasmons superlattice single-particle 27 Iff
330
surface single-particle 158ff electrostatic region 183, 200ff, 204 envelope function 272 EuO 148 evanescent mode 220 exchange interaction 86f, 96 excitons 188f extraordinary wave 265 Fabry-Perot interferometer 23 Faraday direction 239 Fe 140f, 149, 151ff FeF 2 111 ferroelectrics bulk modes 121 displacive 202 hydrogen-bonded 120 pseudo-spin model 120 surface modes 123f, 202 surface reconstruction 122ff ferromagnets bulk magnetostatic modes 130, 133 bulk magnons 88ff, 102, 104, 121, 123, 129, 147f bulk polaritons 236f continuum theory 96ff dipole-exchange modes 137ff double-layer system 136f, 152 films 104, 138, 140ff, 150 Hamiltonians 86f, 120 itinerant-electron 169 superlattice magnetostatic modes 287ff superlattice polaritons 290f superlattice reconstruction 29 Iff surface magnetostatic modes 130ff, 142, 153 surface magnons 93ff, 102, 106ff, 123f, 148 surface polaritons 236ff surface reconstruction 92, 122ff two-interface polaritons 24 Iff see also magnons field-effect transistor 177 Floquet's theorem 4 see also Bloch's theorem fluctuation-dissipation theorem 19, 61, 314ff folded modes acoustic phonon 249ff magnon 298f optic phonon 256ff frequency dispersion, see optical dispersion frequency-energy conversion factors 24 Fresnel factor 59 GaAs 67, 185, 193, 198, 230 see also superlattices: GaP 224, 229, 232f Ge 57, 164
Index grating coupling 219 grating spectrometer 23 gravity waves on liquid 78f Green functions acoustic 50ff, 254 basic properties 31 Iff dipole-exchange 142 jellium 167ff magnetic 99ff, 108 superlattices 254, 262, 268, 283 surface matching 56, 162 guided waves acoustic 50 see also polaritons Hamiltonian electron gas 166 Heisenberg antiferromagnet 109 Heisenberg ferromagnet 86 Ising ferromagnet 87, 120 mean-field for ferromagnet 89 harmonic oscillator 17ff Heisenberg model, see Hamiltonian Hg 81,83 Hooke's law 41 hydrodynamics electron gas 174ff, 203 liquids 74ff inelastic particle scattering 28ff, 68ff InSb 171, 187 intercalated compounds 248 interface modes 137 jellium 166 see also electron gas Kapitza conductance, magnetic 103 KDP (potassium di-hydrogen phosphate) 120 KNiF 3 114 Kronig-Penney model 272 Lamb waves 49 Lame parameters 41, 50 lattice dynamics one-dimensional 34ff superlattice 256ff three-dimensional 40, 69ff leaky modes 141, 205 LEED, see electron diffraction LiF 211 light-emitting tunnel junctions 233ff light scattering, see Brillouin scattering, Raman scattering Lindhard function 166, 169, 181 linear-response theory 16ff, 31, 312ff see also Green functions liquid crystals 248
liquids bulk waves 76 surface waves 77ff long-range surface plasmon (LRSP) 212, 227 Love waves 49ff Lyddane-Sachs-Teller (LST) relation 179 magnetic resonance, see antiferromagnetic resonance, spin wave resonance magnetisation, surface 92 magnetoelastic waves 302f magnetostatic modes antiferromagnets 142ff cylindrical surface 308 ferromagnetic double layer 136f ferromagnets 130ff spherical surface 309 superlattices 287ff magnetostatic region 127 magnons bulk Heisenberg antiferromagnet 1 lOff bulk Heisenberg ferromagnet 88ff bulk Ising, transverse field 121 continuum theory 96ff definition 85 edge modes 307f films 104ff magnon-magnon interactions 304 semi-classical 90 semi-infinite Heisenberg 90ff superlattice 296ff surface acoustic 94 surface antiferromagnetic 113,116 surface Ising, transverse field 124 surface optic 94 MBE, see molecular beam epitaxy mean-field theory 89, 92, 121 metallo-organic chemical vapour-phase deposition (MOCVD) 247 microwave experiments 153f Miller indices 31 mixed excitations 185, 302f see also magnetoelastic waves, polaritons MnF 2 111, 146f, 241 molecular-beam epitaxy 247 monatomic lattice, see lattice dynamics multiple quantum well (MQW) 274 Neel temperature 86 neutron scattering 29 Ni 72,118 nonlinear effects 214ff, 3O3f nonreciprocal propagation 133, 135, 142, 145, 149, 151, 238, 303 opacity broadening 27, 57, 63, 149 optic phonons, see phonons
331
Index optical dispersion 165, 183 ordinary wave 265 Otto configuration, see attenuated total reflection particle scattering, see electron diffraction, inelastic particle scattering, neutron scattering phonons acoustic 34ff, 37, 249ff confined 259 cylinder 309 optic 37f, 73, 178f, 256ff sphere 309 see also elastic waves photoelectron spectroscopy 23, 118 photon-correlation spectroscopy 80f pinning 95 see also surface anisotropy, magnetic plasma frequency 170 plasmons superlattice 276ff superlattice intersubband 282 superlattice intrasubband 281, 286 superlattice surface 280 surface 197, 199 2D electron gas 203f 3D electron gas 157, 170, 175 see also polaritons Pockels tensor, see acousto-optic tensor Poisson's ratio 41 polaritons bulk exciton- 188ff bulk magnon- 23 5f bulk phonon- 184ff bulk plasmon- 184ff cylinder 309 guided-wave 209, 212ff, 230ff, 243, 245f Raman scattering 187, 228ff single-surface 194ff, 229 single-surface, anisotropic media 202f, 245, 270 single-surface, charge-sheet 203f single-surface, electrostatic 200ff single-surface, self-guided 215ff sphere 309 superlattice magnetic 290 superlattice surface 270 surface-exciton 204ff, 225f surface-magnon, antiferromagnetic 240 surface-magnon, ferromagnetic 236ff surface, nonlinear 214ff two-interface 208ff, 230 two-interface magnetic 24 Iff power spectrum 19 Poynting vector 200, 217f p-polarisation, acoustic 44 pseudo-spin model, see ferroelectrics pseudo-surface wave 45
332
Raether-Kretschmann configuration, see attenuated total reflection Raman scattering bulk polaritons 187 folded acoustic modes 253ff folded optic modes 26If principles 23ff superlattice plasmons 284ff surface polaritons 228ff random-phase approximation (RPA) 88, 129, 166, 170, 177, 181 Rayleigh wave 45ff, 64ff, 68, 70, 303, 306 RbMnF 3 114 reciprocal lattice 3, 7 reconstruction, surface 1, 12ff, 92, 122, 124, 291ff reflection acoustic waves 42ff reflectivity, optical 154, 172f, 241 response function 16ff see also Green function, linear-response theory Reststrahl 177ff, 185, 197ff, 267f RHEED, see electron diffraction rough surfaces 234, 309f RPA, see random phase approximation scalar potential, magnetostatic 131 scattering atoms see inelastic particle scattering electrons, see electron diffraction, electron energy loss spectroscopy light, see Brillouin scattering, Raman scattering neutrons, see neutron scattering scattering cross section definition 27 see also differential cross section screening 157 selection rule, polarisation 60 selvedge 12 semiconductor superlattices growth 247 simple band model 27 Iff terminology for electron states 274 see also superlattices SEXAFS 23 SGFM, see Green function Si 57, 65, 68, 163, 172f skin depth, electromagnetic 59 slab modes 48f sound velocity longitudinal 42, 76 transverse 42 space lattice 2, 7 spatial dispersion 165, 170 specific heat, surface acoustic 49
Index magnetic 117, 119 spherical and spheroidal surfaces 307ff spin deviation lOlf spin-flop phase transition 116 spin wave resonance 117 spin waves, see magnons s-polarisation, acoustic 43 sputtering 247 Stokes scattering conservation rules 3If definition 24 magnons 102, 148, 150, 289 via exciton-polaritons 193 see also Brillouin scattering, Raman scattering Stoneley wave 49 stop band 37, 39, 161, 185, 256, 263, 269, 274 strain tensor 41 stress tensor 41,75 superlattices acoustics 249ff Al/W 257 Brillouin scattering 289 cyclotron resonance 274 definitions and terminology 11, 247 dielectric function 267 electron states 271ff ferromagnetic/antiferromagnetic 291 f Fe/Gd 294 GaAs/Al^Gai _xAs 254f, 261f, 267ff, 277, 286 Green functions 262, 268 growth 247 magnetic 287ff magnetic permeability 290f magnetic polaritons 290 magnetic reconstruction 29Iff magnetostatic modes 287ff magnons 296ff
Ni/Mo 289 optical properties 263ff phonons 249ff, 256ff plasmons 276f, 286 Raman scattering 252ff, 257, 261f, 284fif surface modes 256, 269f, 280, 286 surface-active medium 196 surface anisotropy, magnetic 95, 98 surface reconstruction, see reconstruction, surface surface ripple scattering 65fT surface tension 77 SWR, see spin wave resonance TaC 74 tight-binding approximation 160 time-of-flight (TOF) spectra 30, 69f torque equations, magnetic 97, 131, 144 transfer matrix 251f, 258, 264, 272f, 299f twisted state 293f uncertainty principle 62 unit cell 2 UPS, see photoelectron spectroscopy viscosity 75 Voigt configuration
132, 236, 287, 290
water 79, 82 waveguiding condition wedge modes 3O5ff Wigner-Seitz cell 3
214
XPS, see photoelectron spectroscopy Young's modulus 41 yttrium iron garnet (YIG)
117, 153
zero sound 169 ZnO 206f ZnSe 226f
333