PHYSICAL REVIEW B, VOLUME 63, 165429

Weight of zero-loss electrons and sum rules in extrinsic processes that can influence photoemission spectra K. Schulte, M. A. James, P. G. Steeneken, and G. A. Sawatzky Material Science Center, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands

R. Suryanarayanan, G. Dhalenne, and A. Revcolevschi Laboratory de Physico-Chimie de l’e´tat Solide, CNRS, UMR 8648, Baˆt. 414, Universite´ Paris-Sud, 91405 Orsay, France 共Received 16 October 2000; published 5 April 2001兲 It was argued in a recent paper by Joynt 关Science 284, 777 共1999兲兴 that in the case of poorly conducting solids a photoemission spectrum close to the Fermi energy may be strongly influenced by extrinsic loss processes similar to those occurring in high-resolution electron-energy-loss spectroscopy, thereby obscuring information concerning the density of states or one-electron Green’s function sought for. In this paper we present a number of arguments, both theoretical and experimental, that demonstrate that energy-loss processes occurring once the electron is outside the solid, contribute only weakly to the spectrum, and can in most cases be either neglected or treated as a weak structureless background. DOI: 10.1103/PhysRevB.63.165429

PACS number共s兲: 79.60.⫺i, 82.80.Pv, 75.30.Vn

I. INTRODUCTION

Photoemission has been used for years as a reliable technique for probing the electronic structure of occupied states in solids ranging from insulators through semiconductors and metals through superconductors. In his paper Joynt1 provided very convincing and interesting arguments that, especially for badly conducting samples 共roughly ␳ 0 ⭓0.1 m⍀ cm, the Mott value兲, the photoelectron spectrum may be affected by energy-loss structures resulting from interaction with the time-dependent fields set up by the photoelectron receding from the surface of the solid. He argued that the influence of these loss processes can be so strong that the spectrum will be dominated by them, and that therefore the intrinsic information regarding the electronic structure of the solid all but disappears. Since photoemission plays such a prominent role in the discussion of strongly correlated materials like high-T c superconductors, or, more generally, transition-metal oxides, as well as in Kondo and heavy fermion systems, it is of quite some importance to further investigate Joynt’s assertions. In this paper we study Joynt’s arguments, and provide both experimental and theoretical findings that show that the effects due to losses discussed by Joynt are only a small contribution to the total spectrum, and that the zero-energy-loss probability for photoelectrons dominates for samples of either good or bad conductivity. II. INTRINSIC AND PSEUDOINTRINSIC EFFECTS

Compared to other techniques, photoemission provides the most easy and direct measurement of the one-electron Green’s function and the directly related occupied density of states of a solid, if one keeps in mind the possible influence on the photoelectron of a number of ‘‘pseudointrinsic’’ effects.2 By this, we mean, first, the matrix element needed for a description of the amplitude of the optical transition probability from an occupied state to a high-energy unoccupied state in the solid; second, the energy-loss processes occurring on the electron’s path to the surface; and, finally, the 0163-1829/2001/63共16兲/165429共6兲/$20.00

description of the escape of the electron through the surface region into the vacuum. We emphasize that, in the so-called ‘‘sudden approximation,’’ we neglect all interactions and interference effects between the high-energy excited electron and the hole left behind. One also has to contemplate truly intrinsic effects: in a one-electron approximation, the associated one-hole Green’s function is a delta peak at an energy determined by the band dispersion of the occupied states. In reality the electrons in the solid are usually not simple free electrons, but interact with other electrons, phonons, magnons, etc., resulting in one-electron Green’s functions now including a frequencyand momentum-dependent self-energy. For weakly interacting systems the initial delta function spectrum for such an electron broadens 共asymmetrically兲, and attains a frequency distribution for each momentum vector. This basically provides information not only on the quasiparticle dispersion and lifetime, but also on the way the electron is dressed inside the solid, due to the response of its environment to its presence or absence. In strongly interacting systems this selfenergy causes a rather large spreading out of the initial delta peaks describing the one-hole Green’s function, and a description in terms of a quasiparticle with a certain lifetime may break down completely. In these cases it is indeed difficult to separate the intrinsic properties of a one-hole Green’s function from pseudointrinsic effects due to the energy losses suffered by the excited electron on its way out of the solid. These losses are basically dominated by the selfenergy of the excited electron. It is therefore extremely important to find good estimates of these pseudointrinsic contributions to the spectrum, because, in cases where they form a substantial part of the spectrum, one would like to be able to recognize and correct for it and retain only the truly intrinsic phenomena. Experimentally there are several ways of checking the expected influence of these energy-loss processes. The most obvious is to study the energy-loss spectrum of electrons incident on the solid, with an initial kinetic energy equal to that of the escaping photoelectron. These loss spectra pro-

63 165429-1

©2001 The American Physical Society

K. SCHULTE et al.

PHYSICAL REVIEW B 63 165429

vide one with direct information about the self-energy of an excited electron in the unoccupied states of the solid. For high-energy electrons (E k ⭓60 keV), one can think of transmission electron-energy-loss spectroscopy 共EELS兲 measurements. The obtained energy-loss information in these experiments is mainly due to bulk processes. In order to achieve high-energy resolution photoemission spectra, however, measurements are usually performed at low photon energies (E Ph ⬍100 eV) such as He I radiation 共21.22 eV兲, or even less. At these low energies electron-energy-loss processes can be studied with reflection high resolution EELS, but one must then realize that the incident electron hardly penetrates the solid surface. Therefore, the surface loss function is measured rather than the bulk loss function. The bulk loss function is what one would associate with the losses experienced by a photoelectron on its way to the surface. The surface and bulk loss functions, however, are linked to each other, and this means that information about the intrinsic processes can also be retrieved from a reflection experiment, although the relative intensities of certain bulk losses and their surface analogs may differ. Finally, we stress the point that Joynt concentrated on losses occurring after the photoelectron has left the solid, and these are directly related to the reflection EELS features. Another way of obtaining the self-energy information of the excited electron is to study the spectral function in a photoemission measurement of a narrow atomic core level of the solid at a photon energy such that the excited electron will have the required 共low兲 kinetic energy. Although this loss spectrum will be intertwined with the satellite structure due to the self-energy effects involved in the sudden creation of the core hole itself, they will nonetheless provide us with an upper limit on the importance of the energy-loss contributions of the material to its photoemission spectrum. This was suggested and used in Ref. 2 to argue that the broad and intense energy distribution seen in angular-resolved photoelectron spectroscopy of high-T c superconductors was not a result of energy-loss processes, but mainly a direct result of the strongly energy-dependent self-energies in these strongly correlated materials. A more detailed study of this effect in the high-T c superconductors was recently published.3 In the interpretation of spectra, it is thus generally assumed that the photoelectron intensity distribution is a true reflection of the single-particle spectral function, and that the aforementioned pseudointrinsic losses can be identified and reckoned with if necessary.

the case of badly conducting solids and anisotropic materials. He claimed, dependent on the properties of the material under study, that this extrinsic distortion can be so dramatic that instead of observing a sharp Fermi cutoff in the spectrum, the observed spectral distribution will look like that expected for a material with a so-called pseudogap at E F . If correct, this would make photoemission an unsuitable technique for studying the one-electron Green’s function of badly conducting solids such as many of the colossal magnetoresistance materials and the high-T c superconductors, among many others. Joynt substantiated his statement by deriving an expression for the energy-loss probability of an electron once the electron has emerged from the surface. He calculated, in a classical picture, the average energy lost due to the interaction of the time-dependent electric field of the moving electron and the polarizable metal left behind. The response of the metal is approximated by a Drudelike behavior. He then distributed this average energy lost over an energy-loss spectrum using a probability distribution as a weighting function. The basic assumption is that the Born approximation is valid so that each electron suffers at most a single scattering event. This results in

P共 ␻ 兲⫽

2e 2 CL 共 ␻ 兲

␲បv␻2

共1兲

,

in which P( ␻ ) stands for the probability that the electron loses energy ␻ , C is a constant for which Joynt gave a value of ⬇2.57, and L( ␻ ) represents the loss function: L共 ␻ 兲⫽





␻ ⫺1 Im . 4␲ 1⫹ ⑀ 共 ␻ 兲

共2兲

We agree to a large extent with this derivation, except perhaps for the constant C, on which we will focus our attention later in this paper. Let us for the moment assume that the Born approximation and Joynt derivation are correct. We see that the material-related properties enter the equation for the energy-loss distribution via the frequency-dependent dielectric constant of the material. This is not unexpected, since it indeed describes the response of the material to the timedependent field produced by the outgoing electron. Let us, as Joynt did, take as an example the dielectric function ⑀ ( ␻ ) given by the Drude model,

III. EXTRINSIC EFFECTS

We will now focus on extrinsic broadening of photoemission structures such as the Fermi edge. Within photoemission this, up to now, meant just a finite instrumental resolution; however, using the picture that Joynt put forward in his recent paper, we now also have to consider losses suffered by the photoelectron after it has left the solid and is on its way to the detector. These losses are directly comparable to the loss spectrum of a reflection EELS measurement. Joynt argued that these losses, more than anything, will severely distort any sharp feature, such as the Fermi edge, especially in

⑀共 ␻ 兲⫽

4␲i 4␲i␴0 ␴共 ␻ 兲⫽ , ␻ ␻ 共 1⫺i ␻ ␶ 兲

共3兲

and plot P( ␻ ) for different values of the resistivity ␳ 0 ⫽1/␴ 0 and scattering time ␶ 共Fig. 1兲. One sees immediately that, although the weight of P( ␻ ) is distributed differently in each curve, the integral 兰 ⬁0 P( ␻ )d ␻ reaches the same value in the end. There is indeed a well-known general sum rule4 connected to this formula which is independent of material parameters:

165429-2

WEIGHT OF ZERO-LOSS ELECTRONS AND SUM RULES . . .

FIG. 1. Top panel: the loss probability P( ␻ ) calculated for different values of resistivity ␳ and relaxation time ␶ , keeping the surface plasmon energy ␻ S2 P ⫽4 ␲ / 关 ␳ ␶ (1⫹ ⑀ ⬁ ) 兴 fixed at 0.75 eV. Lower panel: demonstration of the sum rule for the loss probability P( ␻ ).

lim

e 2C

k→0 2 ␲

2

បv



⬁d␻

0



Im





⫺1 e 2C ␲ 1 ⫽ . 1⫹ ⑀ 共 ␻ 兲 2 ␲ 2 ប v 2 1⫹ ⑀ ⬁

共4兲

This is a general sum rule that holds for any model of ⑀ ( ␻ ) provided that it is a causal function. It depends only on the velocity of the outgoing photoelectron which is assumed to be constant in the process, consistent with the Born approximation. So, for an electron with a kinetic energy of 20 eV, as used in Joynt’s examples, the integral from zero to infinity gives ⬇0.0843 for the total of the losses. There is of course another sum rule which conserves the integrated intensity of the spectrum. In this sum rule we must include the finite probability P 0 , which represents the probability that an electron suffers no loss at all, so that 1⫽ P 0 ⫹





0

P共 ␻ 兲d␻.

共5兲

Using the results above we find that P 0 ⬇0.957, and therefore the normal single particle density of states 共DOS兲 will dominate the photoemission spectrum. Up to now we have only been concentrating on the freeelectron part of the response function. Of course, there are more contributions, such as phonons and interband transitions, that are contained in the total dielectric function of a real material. The calculations presented by us and by Joynt, which only take into account the Drude model, will therefore always overestimate the influence of the free electron contribution. Let us take a simple example to illustrate this. Our first sum rule pins down the total amount of losses from zero to infinite frequency. This means that other processes, such as phonons, which are not contained in the Drude model, will capture some of the weight carried by the free electron losses

PHYSICAL REVIEW B 63 165429

that we have considered so far. This implies that in a good metal, where the phonon part will be nearly fully screened by the free-electron excitations, the 共low-energy兲 loss spectrum is, to a first approximation, indeed well described by the Drude model. If the material is then changed into a bad conductor, the phonon part of the loss spectrum will gain more and more strength in the low-energy region of the spectrum at the expense of the Drude part because of the sum rule. In both cases however, P 0 will have a fixed value. Furthermore, any phonon contribution will not influence the photoemission spectroscopy 共PES兲 spectrum, on the low-energy scale around the Fermi energy, in a smooth way, but rather produce a step in the convoluted photoemission-EELS spectrum, since these phonon losses are peaked at the phonon frequency and never overdamped. The same holds for a good metal, where there is a clearly defined loss peak at the plasma frequency. Thus, we can conclude that at 20 eV, 兰 ⬁0 P( ␻ )d ␻ ⬇0.0843 represents an upper limit to the losses due to the Drude part. This very general result is in direct conflict with the assumption made by Joynt that P 0 is probably very small. He arrived at this assumption because he did not use the general dielectric function sum rule 关Eq. 共4兲兴, and he therefore concluded that processes other than the freeelectron losses under consideration will reduce P 0 further, but with an amount that is unknown. This led him to take P 0 to be a fit parameter, and, in order to fit the data, he used values close to zero for it. In our opinion it is not possible to make an independent fit parameter out of P 0 , as Joynt did, since its value is fixed by the sum rules and it is, in essence, independent of material constants. From an experimental point of view, we know from reflection EELS experiments 共see, e.g., Fig. 4兲, at incoming energies of around 20 eV, that the elastic peak 共which represents P 0 ) is by no means close to zero for any material. Only for very low incoming energies 共below ⬇10 eV), or when special surface waveguidelike conditions are met,5 can the zero-loss peak be strongly suppressed. Besides this, in reflection EELS the loss probability P( ␻ ) is even twice as strong as in this photoemission scenario, since the electron can lose energy both on the incoming and outgoing trajectories, but otherwise the effect described by Joynt is the same. One can therefore also use the method presented by Ibach and Mills,6 for calculating reflection EELS spectra, in the case of this photoemission problem. It is interesting to note that one obtains the same result except for a different numerical prefactor 共a detailed derivation can be found in Ref. 7兲. This presumably has its origin in a difference in Fourier transform convention and, in Mills’s case,7 avoiding integrals such as Eq. 共1兲 in Joynt’s paper, which is not readily solvable analytically. Using the prefactor obtained by Mills, the equation for the losses reads P共 ␻ 兲⫽





e2 ⫺1 Im . ប␻v 1⫹ ⑀ 共 ␻ 兲

共6兲

When we use this form of P( ␻ ) in our sum rule, it turns out that the losses are in fact much more severe than with Joynt’s original prefactor. We now have 兰 ⬁0 P( ␻ )d ␻ ⬇0.65 共again at a kinetic energy of the electron of 20 eV兲, leaving P 0 at 0.35.

165429-3

K. SCHULTE et al.

PHYSICAL REVIEW B 63 165429

FIG. 3. The resistivity of our La1.2Sr1.8Mn2 O7 samples in the ab plane 共circles兲 and the c direction 共squares兲 as a function of temperature 共from Ref. 9兲. FIG. 2. Comparison between Joynt’s original calculation with P 0 ⫽0.01 and 兰 P( ␻ )d ␻ ⫽0.0843 共lower panel兲, and our calculation using Mills’ prefactor for the loss probability P( ␻ ) and obeying the sum rule P 0 ⫽1⫺ 兰 P( ␻ )d ␻ ⫽0.35 共top panel兲. The parameters are specified in the figure, and T⫽38 K. Although the shape of the spectrum is affected, there is no clear sign of a pseudogap for the bad metal in our case.

However, the zero-loss part will still be large enough to dominate the Fermi cutoff, if one thinks again in terms of the complete dielectric function being involved, instead of just the Drude part. On the other hand, this result does imply that working in the Born approximation is no longer valid, and a strong interaction picture, also containing multiple losses, needs to be applied. Unfortunately, this is less straightforward to derive for a continuous spectrum of excitations. In the case of discrete, well-defined plasmon losses due to core-level excitations, the loss spectrum was calculated by Langreth.8 His conclusion was that multiple plasmon losses are seen distributing the total energy loss over a wider energy range, at the expense of the low-energy losses. On the basis of this calculation we argue that multiple-scattering corrections will not strongly influence our arguments, since they will redistribute the weight of the losses over a wider energy range and thereby reduce the relative influence in the low-energy loss region. If we, for the moment, stick to the single scattering scenario and take for the losses only the Drude contribution, we can calculate the photoemission spectrum for the same material parameters as Joynt used, except taking P 0 ⫽0.35. This is depicted in Fig. 2. The lower panel contains Joynt’s original calculation,using P 0 ⫽0.01. In this picture we see that, although the line shape of the photoemission spectrum is affected by the losses in the case of the bad metal 共where the Drude contribution is surely overestimated兲, there is still finite weight at the Fermi energy, and therefore the effect does not create a clear pseudogap structure. IV. EXPERIMENT

As a last discussion point, we will present the case of La1.2Sr1.8Mn2 O7 , a double layered colossal magnetoresis-

tance oxide with a ferromagnetic metal 共at low temperature兲 to a paramagnetic insulator transition at 125 K. It was grown by means of the traveling solvent floating-zone method.9 This material is a good candidate for testing Joynt’s assumption that materials with high resistivity will be more prone to the influence of losses on the Fermi region in photoemission spectra than those with low resistivity, as the resistivity changes by roughly two orders of magnitude from below to just above the phase transition 共see Fig. 3兲.9 In order to have all the ingredients to test Joynt’s hypothesis, we performed both reflection EELS at 20.5-eV incoming electron energy and angle-integrated PES using a He I source. For both measurements the sample was cleaved in situ at a temperature of 60 K, and at a base pressure of 8 ⫻10⫺11 in the case of EELS and 4⫻10⫺11 in the case of PES. For the EELS measurement we used a customized VG electron monochromator/analyzer system. The analyzer is rotatable since it is mounted on a two-axis goniometer. The samples are held by a Janis ultrahigh-vacuum flow cryostat. The incoming/outgoing angle of the electron beam was 35° with respect to the surface normal. The incoming electron energy was 20.5 eV. The total zero loss peak full width at half maximum is 60 meV. The acceptance angle of the analyzer is 2°. For ultraviolet photoemission spectroscopy the system consists of an Omicron helium discharge lamp, a VG Clam 2 electron analyzer 共acceptance angle 8°, overall energy resolution 50 meV兲, and again a Janis cryostat. As these samples deteriorate even at these low pressures in a matter of hours, we ensured that measurements were performed within 2 h after the cleave. That is before the peak at 9-eV binding energy in the PES spectra starts to appear, which is associated with a change in oxygen stoichiometry at the surface.11 The regions around E F of the satellite and background noise-corrected photoemission spectra are shown in Fig. 5 for T⫽60 K 共solid line兲, 140 K 共dashed line兲 and 180 K 共dotted line兲. The inset depicts the full spectrum, taken at 60 K, before and after the temperature cycle. Since we performed EELS at a finite incoming angle with respect to the surface normal, ( ␪ in ⫽35°), we have to apply a correction factor, as described by Ibach and Mills,6 to extract the surface loss function Im关 ⫺1/1⫹ ⑀ ( ␻ ) 兴 . We then

165429-4

WEIGHT OF ZERO-LOSS ELECTRONS AND SUM RULES . . .

FIG. 4. Top panel: reflection EELS spectra, as taken, of La1.2Sr1.8Mn2 O7 , at 50 K 共solid line兲 and 150 K 共dotted line兲. The filled black curve is the 50-K spectrum divided by 150 to show the relative weakness of the loss features with respect to the zero-loss electrons. Lower panel: geometry-corrected spectra, according to Ref. 6 multiplied by e 2 /ប ␻ v to obtain P( ␻ ). Inset: first 400 meV, showing the stronger presence of a surface phonon in the insulating regime, relative to the metallic phase.

multiply this by e 2 /ប ␻ v to obtain P( ␻ ), as described in Eq. 共6兲, to be able to use for our calculation of a PES spectrum. In Fig. 4, top panel, we show the EELS spectra, as taken, at 50 K 共solid line兲 and 150 K 共dashed line兲, including the zero-loss line. In the lower panel we show P( ␻ ) calculated from the data after subtraction of the zero-loss line. We then take this P( ␻ ) to calculate its influence on a PES spectrum, assuming a constant density of states, using various values of P 0 . This is depicted in the lower panel of Fig. 5. It can be seen from this figure that we cannot reproduce the photoemission spectrum at all, using Joynt’s idea, unless P 0 ⫽0. In other words, for any finite value of P 0 combined with a finite DOS at the Fermi energy, one will obtain a finite Fermi cutoff. To obtain a rough estimate of P 0 from our EELS experiment, we can integrate the zero-loss peak separately, and compare it to half the integral of the entire loss region up to 10 eV as measured in our EELS spectra. Of course, we have to keep in mind that we use our detector in a mode which selects electrons within an opening angle of 2° around the specular reflection, and it is therefore not a fully angleintegrated measurement, which makes us underestimate the losses relative to the zero-loss probability. However, if we proceed in this way, for both 50 and 150 K we obtain a ratio of P 0 : P( ␻ )⫽0.82:0.18. This at least shows that P 0 is not close to zero, and P 0 does not depend on temperature 共or resistivity兲 of our sample. These findings agree with angleresolved PES measurements performed by Dessau and

PHYSICAL REVIEW B 63 165429

FIG. 5. Top panel: PES spectra at different temperatures above and below the phase transition in La1.2Sr1.8Mn2 O7 close to the ‘‘Fermi energy’’ the inset shows the spectra up to 12-eV binding energy, before and after the temperature cycle. Lower panel: calculated PES spectra using the loss probability P( ␻ ) obtained from the EELS data of Fig. 4, and assuming a constant DOS, for various values of the zero-loss probability P 0 as indicated in the figure.

Saitoh10 and Saitoh et al.11 In their papers, they use the fact that Joynt expected the loss effect to be angle independent to show that the angle, at which the smallest change is observed in going from above to below T c , is indicative of the maximum magnitude of the effect; this, in their experiments, turned out to be negligible. V. CONCLUSIONS

In conclusion, we have argued that Joynt indeed raised an important question regarding the influences of extrinsic losses on low-energy photoelectrons, but we disagree with his statement that the losses will take the upper hand in determining the shape of the spectrum around the Fermi energy. This is an immediate consequence of the presented loss function sum rule, that renders the zero loss probability substantially finite. Joynt’s effect may become a point of concern, however, when very low photon energies are used. However, down to photon energies such as the often used He I line 共21.22 eV兲, we have demonstrated, both experimentally and theoretically, that this is not the case. Of course, since the losses, resulting from Mills’ derivation, are by no means a small perturbation in this classical, single scattering approach, a full quantum-mechanical treatment including multiple scattering is called for. Furthermore, we are not able to reproduce Joynt’s formula exactly as far as the prefactor is concerned, instead we agree with the treatment by Mills,7 which is self-consistent and avoids integrals with questionable convergence. We also have shown in the case of La1.2Sr1.8Mn2 O7 that we cannot reproduce the photoemission spectrum using a finite density of states up to the Fermi energy together with a finite value for P 0 , so that we must

165429-5

K. SCHULTE et al.

PHYSICAL REVIEW B 63 165429

We would like to thank D. L. Mills for letting us use his calculation of P( ␻ ) for our simulations and the stimulating

discussions, and M. Mostovoy and L. H. Tjeng for their valuable contributions along the way. This research was supported by the Netherlands Foundation for Fundamental Research on Matter 共FOM兲 with financial support from the Netherlands Organization for the Advancement of Pure Research 共NWO兲. The research of M.A.J. was supported through a grant from the Oxsen Network.

R. Joynt, Science 284, 777 共1999兲. G. A. Sawatzky, in Proceedings of High T c Superconductivity Symposium, Los Alamos, 1989, edited by K. S. Bedell, D. Coffey, D. E. Meltzer, D. Pines, and J. R. Schrieffer 共AddisonWesley, Reading, MA, 1990兲, p. 297. 3 M. R. Norman, M. Randeria, H. Ding, and J. C. Campuzano, Phys. Rev. B 59, 11 191 共1999兲. 4 G. D. Mahan, Many Particle Physics 共Plenum, New York, 1990兲, p. 467. 5 J. J. M. Pothuizen, Ph.D. Thesis, Rijksuniversitteit Groningen, 1998. 6 H. Ibach and D. L. Mills, Electron Energy Loss Spectroscopy and Surface Vibrations 共Academic, New York, 1982兲, Chap. 3.

D. L. Mills, Phys. Rev. B 62, 11 197 共2000兲. In the end, Mills claims that the classical description of the energy, lost by an electron retreating from the surface, should be interpreted in terms of a shift in energy of all emitted electrons and not in terms of a loss spectrum. We agree, in this regard, with Joynt: the average energy calculated classically, would, in a quantummechanical treatment, result in an energy-loss spectrum rather than a rigid energy shift. 8 D. C. Langreth, Phys. Rev. Lett. 26, 1229 共1971兲. 9 W. Prellier et al., Physica B 259-261, 833 共1999兲. 10 D. S. Dessau and T. Saitoh, Science 287, 767a 共2000兲; R. Joynt, ibid. 287, 767a 共2000兲. 11 T. Saitoh et al., Phys. Rev. B 62, 1039 共2000兲.

conclude that there is a true pseudogap in this material both below and above the phase transition. ACKNOWLEDGMENTS

1 2

7

165429-6

Weight of zero-loss electrons and sum rules in extrinsic ...

Apr 5, 2001 - to recognize and correct for it and retain only the truly in- trinsic phenomena. Experimentally there are several ways of checking the expected influence of these energy-loss processes. The most obvious is to study the energy-loss spectrum of electrons incident on the solid, with an initial kinetic energy equal ...

84KB Sizes 0 Downloads 27 Views

Recommend Documents

Cyclotron Resonance of Electrons and Holes in ... - APS Link Manager
Apr 2, 2015 - (Received December 16, 195O). An experimental and theoretical discussion is given of the results of cyclotron resonance experiments on charge carriers in silicon and germanium single crystals near O'K. A description is given of the ligh

Arrangement of Electrons in Atoms - TeacherWeb
Arrangement of Electrons in Atoms. MIXED REVIEW. SHORT ANSWER Answer the following questions in the space provided. 1. Under what conditions is a ...

Rammer, Quantum Transport Theory of Electrons in Solids, A Single ...
Rammer, Quantum Transport Theory of Electrons in Solids, A Single Particle Approach.pdf. Rammer, Quantum Transport Theory of Electrons in Solids, A Single ...

Electrons, holes, and spin in Nd2ÀxCexCuO4À
Mar 7, 2003 - of the valence band in the undoped compound showing that the parent compound of this electron doped superconductor is similar to the parent ...

Thermal characterization of intrinsic and extrinsic InP ...
Apr 2, 2003 - attention in the context of fabrication of electronic and optoelectronic devices .... Introduction of a foreign atom into the host lattice creates more ...

Unravelling extrinsic and intrinsic factors of the early ...
comprehensive database including all records of blastozoans was built to provide quantitative analyses of .... regional databases was tested by the comparison of their rarefied .... test (bottom triangular matrix) and Kendall correlation (top triangu

Sum of us
Page 1 of 23. Windows with activator.Tomb raider 1996 PC.95784035019 - Download Sumof us.Star wars rebels 1080p s01.Italso shows who areliving in. theflatsand thestaff, which areallocated to theseflats. The main role ofthisentity is to providethescho

Annuitiy and Lump-Sum Decisions in Defined ... - Semantic Scholar
directly with the degree to which plan rules restrict the ability to choose a partial or ...... Journal of Pension Economics and Finance, 3(2): 109–43. ... labor unions; health care providers and insurers; government organizations; and service firm

THE VALUE OF ZERO-SUM STOPPING GAMES IN ...
every given stage only one of the players is allowed to stop; the identity of that player ... ‡MEDS Department, Kellogg School of Management, Northwestern University, Evanston, IL. 60208, and .... Define a randomized stopping time φ by.

Annuitiy and Lump-Sum Decisions in Defined ... - Semantic Scholar
calculate an annuitization rate of 61 percent, meaning that 61 percent of ..... risk the loss of their investments in view of their remaining lifespans. ..... EBRI's Blog supplements our regular publications, offering commentary on questions received

Shoemaker's Key, Extrinsic Dispositions, and Causal ...
Jul 16, 2009 - In this case, I may even be able to use a simple off-the-shelf model from classical mechanics such ..... epistemic representation—the similarity account and the structural account. ..... sure the toboggan won't go too fast. So, for .

The relationship within and between the extrinsic and intrinsic systems ...
Apr 1, 2007 - 360, 1001–1013. Binder, J.R., Frost, J.A., Hammeke, T.A., ... other: a social cognitive neuroscience view. Trends Cogn. Sci. 7,. 527–533.

Limited Feedback and Sum-Rate Enhancement
Nov 3, 2012 - obtained by considering the effect of feedback overhead on the total throughput of the MIMO IMAC model. I. INTRODUCTION. Interference ...

The relationship within and between the extrinsic and intrinsic systems ...
Apr 1, 2007 - aDepartment of Biomedical Engineering, School of Computer and Information .... system. These resting state functional network patterns have been .... peaks in a previous study (Tian et al., in press), and the one for the.

intrinsic and extrinsic motivators to study industrial ...
proper focus group and make it comfortable for their students to express their feelings without .... I wanted to be a teacher but teachers do not earn money. V.

Statuts - Electrons Libres.pdf
M. Benoît Marchand, né à Namur le 02/02/1964 et domicilié rue du village 54 à 5352 Perwez-Haillot,. M. Szaniszló Szöke, né à Uccle le 06 novembre 1963 et ...

Changes in Rules and Regulations of APTO.PDF
Page 1 of 2. Islamic Republic of Afghanistan. Ministry of Commerce and Industries. Directorate of Central Business Registry and Intellectual Property. No. 114 Dated: 03.04.2016 – 15.01.1395. Suggestions Sentences/Order. Observed. 03.04.2016 – 15.

Transformations of kinetic energy of free electrons into ...
The distance is measurable as soon as the number of atoms per unit ... however, which has the special quality of consisting of electrically charged particles and .... appears in the spectrogram next to a continuous spectrum in the long-wave.

Stress and gradient weight in Portuguese
(MaxEnt) grammar [3, 13, 6]. Crucially, the vast majority of words in the lexicon are accounted for (≈ 80%) by one single predictor, namely, weight—including a portion of irregular words, where stress is unpredictable. Unlike previous analyses, n

Transformations of kinetic energy of free electrons into ...
Transformations of kinetic energy of free electrons into excitation energy of atoms by impacts. Nobel Lecture, December11 , 1926. Ladies and gentlemen!

Guideline on clinical evaluation of medicinal products used in weight ...
Jun 23, 2016 - The scope of this guideline is restricted to the development of pharmacological ... Alternative analyses based on responder definitions that also ...

Weight Loss in the Elderly
diagnostic testing that may yield few useful data if the results of these initial tests are normal. The first steps .... intestinal systems. Anthropometric measurements ...