Are the surfaces of CrO2 metallic? - CiteSeerX

32 downloads 146 Views 563KB Size Report
Jul 3, 2007 - (Some figures in this article are in colour only in the electronic version) ..... A SP-UPS study of epitaxial CrO2(100) films grown on TiO2(100) by Dedkov et al ..... and the valence features after sputtering is a signature of sample ...
IOP PUBLISHING

JOURNAL OF PHYSICS: CONDENSED MATTER

J. Phys.: Condens. Matter 19 (2007) 315207 (18pp)

doi:10.1088/0953-8984/19/31/315207

Are the surfaces of CrO2 metallic? C A Ventrice Jr1,9 , D R Borst2 , H Geisler3 , J van Ek4 , Y B Losovyj5 , P S Robbert2 , U Diebold6 , J A Rodriguez7 , G X Miao8 and A Gupta8 1

Department of Physics, Texas State University, San Marcos, TX 78666, USA Department of Physics, University of New Orleans, New Orleans, LA 70148, USA 3 Institute for Environmental and Industrial Science, Texas State University, San Marcos, TX 78666, USA 4 Seagate Technology, Bloomington, MN 55435, USA 5 Center for Advanced Microstructures and Devices, Louisiana State University, Baton Rouge, LA 70806, USA 6 Department of Physics, Tulane University, New Orleans, LA 70118, USA 7 Department of Chemistry, Brookhaven National Laboratory 11973, USA 8 Center for Materials for Information Technology, University of Alabama, Tuscaloosa, AL 35487, USA 2

E-mail: [email protected]

Received 20 November 2006, in final form 28 January 2007 Published 3 July 2007 Online at stacks.iop.org/JPhysCM/19/315207 Abstract Previous photoelectron spectroscopy studies of CrO2 have found either no density of states or a very low density of states at the Fermi level, suggesting that CrO2 is a semiconductor or a semi-metal. This is in contradiction to calculations that predict that CrO2 should be a half-metallic ferromagnet. Recently, techniques have been developed to grow high-quality epitaxial films of CrO2 on TiO2 substrates by chemical vapour deposition. We present photoelectron spectroscopy measurements of epitaxial CrO2 (110)/TiO2 (110) and CrO2 (100)/TiO2 (100) grown using a CrO3 precursor. In addition, measurements of epitaxial Cr2 O3 (0001)/Pt(111) films grown by thermal evaporation of Cr in an oxygen atmosphere are presented as a reference for reduced CrO2 films. The measurements of the CrO2 surfaces show no emission at the Fermi level after sputtering and annealing the surfaces in oxygen, even though our soft core photoemission data and low-energy electron diffraction measurements provide evidence that stoichiometric CrO2 is present. The consequence of this is that neither surface of CrO2 is metallic. This behaviour could result from a metal to semiconductor transition at the (110) and (100) surfaces. (Some figures in this article are in colour only in the electronic version)

9 Author to whom any correspondence should be addressed.

0953-8984/07/315207+18$30.00 © 2007 IOP Publishing Ltd

Printed in the UK

1

J. Phys.: Condens. Matter 19 (2007) 315207

C A Ventrice Jr et al

1. Introduction Chromium dioxide (CrO2 ) is a unique ferromagnetic oxide that is predicted to be a half-metallic ferromagnet [1]. Half-metallic ferromagnets are conducting solids whose conduction electrons undergo magnetic ordering with a spin polarization of 100% at 0 K [2]. This unique property occurs when there are charge carriers of only one spin orientation at the Fermi level E F or when both spin orientations are present but the carriers of one spin orientation are itinerate, whereas the carriers of the opposite spin orientation are localized [3, 4]. The spin polarization of a metallic ferromagnetic material can be defined as

P=

N↑ − N↓ , N↑ + N↓

(1)

where N ↑ and N ↓ are the density of states of the spin-up and spin-down electrons at E F , respectively [5]. Most conventional ferromagnetic materials have a spin polarization of less than 50%. For instance, the spin polarizations of Fe, Co, and Ni measured by superconducting tunnelling spectroscopy give values of 40%, 35%, and 23%, respectively [6]. These elemental ferromagnets have either a partially or fully spin-polarized 3d band. However, the unpolarized 4s band also crosses the Fermi level and contributes enough to the density of states to reduce the spin polarization below 50%. There are several materials that are predicted to be half-metallic ferromagnets. Most of these materials are metal oxides and include chromium dioxide (CrO2 ) [1], magnetite (Fe3 O4 ) [7], the mixed valence magnetites (La1−x Ax MnO3 ; A = Ca, Ba, Sr; x ∼ 0.3) [8], and the double perovskites (Sr2 FeAO6 ; A = Mo, Re) [9]. Of these, chromium dioxide, which is isostructural with SnO2 and the rutile form of TiO2 , has the simplest crystal structure and has probably been the most thoroughly studied with respect to its predicted half-metallic property [3]. The primary industrial application of chromium dioxide is as a magnetic recording medium for video and audio tapes. Its widespread use in magnetic recording is not a result of its half-metallic property but because it can be grown as a powder composed of needle-like crystallites, which makes it relatively easy to magnetize, and because its Curie temperature is well above room temperature (TC ∼ 390 K) [10]. The spin-resolved density of states of CrO2 calculated using the local spin density approximation (LSDA) of density functional theory (DFT) is shown in figure 1. The details of these calculations are given in a previous publication [11]. As seen in figure 1, the Cr 3d band is exchange split by approximately 2 eV, leaving the majority band partially filled and the minority band completely empty. Using the convention of Coey and Venkatesan [2], CrO2 is a type IA half-metal. Half-metallic materials that are predicted to have a completely filled majority band and a partially filled minority band are type IB half-metals. The calculated spinresolved density of states presented in figure 1 agrees qualitatively with the results of several other groups that also predict a spin-split band structure of a type IA half-metal for CrO2 using either conventional LSDA [1, 11–14], LSDA with the inclusion of a Hubbard parameter U to account for on site Coulombic interactions (LSDA + U ) [13, 14], or LSDA with dynamical mean field theory (LSDA + DMFT) [15]. The primary interest in half-metallic ferromagnetic materials is for the development of magnetic sensors and devices with an enhanced performance over those using conventional ferromagnetic materials. One example is the giant magnetoresistance (GMR) spin valve [16], which is a device that consists of two ferromagnetic layers that are separated by a nonmagnetic spacer layer as shown in figure 2. One of the ferromagnetic layers is usually grown on an antiferromagnetic pinning layer, which makes it insensitive to moderate magnetic fields (i.e. a magnetically hard layer). The other layer is usually separated from the first with a conducting 2

J. Phys.: Condens. Matter 19 (2007) 315207

C A Ventrice Jr et al

Density of States

10

Majority Spins

0

Minority Spins

-10 10

0

Binding Energy (eV) Figure 1. Calculated spin-resolved density of states per eV per formula unit for CrO2 . The majority/minority spin channel is plotted as a positive/negative density of states value.

nonmagnetic layer. This results in a ferromagnetic layer that can switch its magnetization direction with the application of relatively small fields (i.e. a magnetically soft layer). If a current is passed through the device either parallel to the plane (current in plane (CIP) mode) or perpendicular to the plane (current perpendicular to plane (CPP) mode) there will be a change in resistance as the magnetically free layer switches its magnetic orientation relative to the pinned layer. In other words, the electron will normally have a higher probability of scattering as it passes from one magnetic layer to the other in the antiparallel state since it sees a lower density of states in the magnetic layer of opposite polarity. In GMR devices based on conventional ferromagnetic materials such as permalloy, the change in resistance is usually less than 20% at room temperature. If two half-metallic ferromagnetic layers are used instead, changes in resistance of 50% or larger should be possible upon switching from a parallel to an anti-parallel magnetic orientation because of the absence of minority states in the half-metallic ferromagnetic layers. The development of tunnelling magnetosresistance (TMR) devices based on half-metallic ferromagnetic materials has also received quite a bit of attention recently [17–22]. The main applications of TMR devices are for magnetic field sensors and for magnetoresistive random access memory (MRAM). The TMR and GMR devices have a similar construction, except that the nonmagnetic spacer layer is replaced with an insulating barrier that is thin enough (∼2 nm) for a measurable quantum mechanical tunnelling current to be detected. A schematic of a TMR device is shown in figure 3. When a bias is applied between the two ferromagnetic layers, electrons will tunnel from one layer to the other through the insulating intralayer. As the free magnetic layer switches from the parallel to the antiparallel magnetization orientation, there will be a drop in the tunnelling current since the electrons are now being injected into minority states instead of majority states. A simple model for spin polarized tunnelling which neglects spin-flip scattering at the interfaces or within the insulating layer was developed by Julli´ere in 1975 [23] and is given by TMR =

2 P1 P2 . 1 − P1 P2

(2)

This model relates the TMR effect to the spin polarization P in each ferromagnetic layer. Using conventional ferromagnetic layers with an AlOx barrier, TMR of ∼50% has been recorded [24]. If the conventional ferromagnetic layers are replaced with half-metallic ferromagnetic layers, the tunnelling current should go to zero as the two layers switch from a parallel to an antiparallel magnetic orientation (i.e. an infinite TMR effect) if there is no spin-flip scattering in the films 3

J. Phys.: Condens. Matter 19 (2007) 315207

C A Ventrice Jr et al

Free Layer

FM

FM

FM

FM

AF

AF

Nonmagnetic Spacer Pinned Layer

Anti-parallel Layers

Parallel Layers

Figure 2. Schematic of a GMR device. As majority-spin electrons pass from one ferromagnetic material to the other, they will normally have a higher probability of scattering in the anti-parallel configuration since the transport is into the minority states of the second ferromagnet. Free Layer

FM

FM

FM

FM

AF

AF

Insulating Spacer Pinned Layer

Parallel Layers

Antiparallel Layers

Figure 3. Schematic of a TMR device. For two ferromagnetic electrodes with a parallel magnetic alignment, majority-spin electrons can access majority-spin states on the other side of the insulating spacer, resulting in a tunnel current if a bias V is applied between the electrodes. If the magnetic orientation of the free layer is switched to the antiparallel alignment, a drop in tunnel current is expected to occur since the density of states available for majority-spin transport will be reduced.

or at the interfaces. This predicted very low quiescent current in the antiparallel state is the primary reason for the interest in using half-metallic materials in MRAM applications. Although the performance of devices based on half-metallic ferromagnetic materials is predicted to be superior to those based on conventional ferromagnetic materials, in almost all published studies where one or more of the ferromagnetic electrodes was replaced with a half-metallic ferromagnet, the performance was degraded instead of enhanced [17–22, 25–27]. Various reasons have been given for the poor performance of devices that are based on halfmetallic ferromagnetic materials. One of the most obvious reasons is spin-flip scattering during the transport process, which can be caused by interfacial roughness or from disorder within the ferromagnetic electrodes or in the nonmagnetic spacer layers. The stoichiometry of the half-metallic material at its surface or interface can also be an issue. For instance, the rutile structured CrO2 will reduce to the corundum structured Cr2 O3 at temperatures above ∼400 ◦ C at atmospheric pressure [28]. Therefore, it is generally accepted that a Cr2 O3 surface layer can form under vacuum processing conditions at elevated temperatures [29–33]. Ideally, devices based on heteroepitaxial layers with a low lattice mismatch should be largely free of nonstoichiometries and have a relatively low defect density. An advancement towards this goal was recently achieved by Miao et al [27]. In this study, heteroepitaxial bilayers of CrO2 /SnO2 were grown on TiO2 (100) substrates by chemical vapour deposition (CVD). Deposition of Co on the insulating SnO2 layers resulted in the formation of TMR devices after patterning. The resistance as a function of applied magnetic field at 10 K for a device with a 1.7 nm SnO2 barrier gave the maximum TMR value of only 14%. By assuming an effective spin polarization of 35% for the Co overlayer, the maximum observed TMR value results in a spin polarization of only 19% for the CrO2 layer [27]. Since the actual performance of devices based on materials that are predicted to be halfmetallic ferromagnets is almost always worse than those based on conventional ferromagnetic 4

J. Phys.: Condens. Matter 19 (2007) 315207

C A Ventrice Jr et al

materials, one must ask the question: Is there any direct experimental evidence of halfmetallicity for any of the materials predicted to be half-metallic ferromagnets? To be a halfmetallic ferromagnet, a material must be a metal and have and a spin polarization of 100% at the Fermi level. In this article, the issue of whether or not the surfaces of CrO2 are half-metallic is addressed. Methods of measuring this property will be described, and a review of previous experimental results and our recent photoelectron spectroscopy results on epitaxial CrO2 films will be presented. 2. Measurement of half-metallicity in ferromagnetic materials Ideally, one would like to perform an experiment where the density of states and spin polarization at the Fermi level ( E F ) could be measured simultaneously with the electrical conductivity of the material to confirm directly that it is a half-metallic ferromagnetic material. In practice, the electrical conductivity is measured separately from the spin polarization and the density of states at E F . Experimental techniques that can be used to determine the spin polarization include spin-resolved ultra-violet photoelectron spectroscopy (SP-UPS) [34] and various transport measurement experiments using point contacts or tunnelling junctions either between two ferromagnetic electrodes or one ferromagnetic and one superconducting electrode [3]. To determine if a material is a metal, a combination of experimental techniques is needed. Ultra-violet photoelectron spectroscopy (UPS) measurements can be used to determine the occupied density of states [34, 35], inverse-photoelectron spectroscopy measurements can be used to determine the unoccupied density of states [34], and temperature-dependent resistivity measurements show whether the material is a conductor or an insulator [36]. For instance, if the resistivity of a material decreases as its temperature approaches 0 K, the material is a conductor, but this does not uniquely determine whether the material is a metal or a semi-metal. In semi-metals, the density of states of an energy band just crosses E F either from the conduction band or valence band side [37]. This results in a resistivity that decreases with decreasing temperature; however, the low density of states at E F for semimetals results in resistivities that are typically an order of magnitude or more higher than for metals. To determine whether a material with a low conductivity is just a metal with a high defect density or a semi-metal, both photoelectron spectroscopy and inverse-photoelectron spectroscopy measurements are needed to determine the density of states both below and above EF. 2.1. Electrical conductivity Early measurements of the electrical conductivity of CrO2 provide conflicting values, probably due to differences in sample purity and because these measurements were performed on compacted powders [28]. Although large single crystals of CrO2 are not available, techniques have been developed to grow high-quality epitaxial films on TiO2 substrates by CVD [38–42]. Measurements of the resistivity of CrO2 epitaxial films as a function of temperature show a continuous drop in resistivity as the temperature approaches 0 K, which indicates that CrO2 is a conductor [38–44]. For instance, Gupta et al [44] have measured a room temperature resistivity of 230 μ cm that drops to 2 μ cm at 5 K for transport along the c-axis of epitaxial CrO2 (100) films grown on TiO2 (100). A comparison of resistivities of Cu (an s-metal), Fe (a ferromagnetic d-metal with partial s character), Bi (a semi-metal), and CrO2 at 273 and 77 K are shown in table 1. At 273 K, CrO2 is a rather poor conductor. This is a general characteristic of most conducting oxides and is attributed to the large cross section for scattering of conduction electrons with optical-phonons and other collective excitations in the oxide. At 77 K, the resistivity of CrO2 drops to about a fifth of the resistivity of Bi, but it is still an order of 5

J. Phys.: Condens. Matter 19 (2007) 315207

C A Ventrice Jr et al

Table 1. Resistivities in μ cm of various conductors at 273 and 77 K.

Cu [45] Fe [45] Bi [45] CrO2 [44]

ρ (273 K)

ρ (77 K)

1.6 8.9 107 200

0.2 0.7 35 7

magnitude greater than that of Fe at that temperature. The resistivity values in table 1 for CrO2 are for epitaxial films. Because of lattice mismatch with the TiO2 substrate, dislocations will be present in the films; therefore, the measured value of 7 μ cm is probably an upper limit to CrO2 ’s resistivity at 77 K. 2.2. Spin polarization measured by Andreev reflection Interpreting spin polarization values measured by transport through a magnetic tunnel junction or across a superconductor/ferromagnetic interface is complicated by scattering processes that can occur at the interfaces or the intralayers of the junction [4]. In addition, point contact techniques such as Andreev reflection [46], which is performed by making direct contact between a superconducting tip and the surface of the substrate, can result in damage to the crystal structure at the tip–surface interface and may affect the local electronic and magnetic structure. For some transport measurement techniques, an enhanced polarization may be measured if multiple reflections occur within the barrier. Spin polarizations of CrO2 using the point-contact Andreev reflection technique have ranged from 81% [41] to 98% [42]. Variations in the measured values of the spin polarization using this technique probably depend on the sample growth techniques and differences in the sample–tip interaction. It is important to note that this technique provides a spin polarization of electrons within a few meV of the Fermi level, which is the energy range that governs the transport properties in devices. However, this technique provides little information about the magnitude of the density of states at E F , which also affects the material’s transport properties. 2.3. Spin-polarized photoelectron spectroscopy The most direct measurement of spin polarization is from SP-UPS. Photoemission spectra can be measured with photons from either a gas discharge lamp or a synchrotron light source and are almost always performed under ultra-high vacuum (UHV) conditions. This technique is a photon-in/electron-out process. The kinetic energy KE of the photoelectrons excited by an incident photon of energy hν is given by KE = hν − eφ − E B ,

(3)

where φ is the work function of the spectrometer and E B is the binding energy of the electron measured with respect to E F [35]. By placing a Mott spin polarimeter at the collector of the electron spectrometer, spin-resolved photoemission spectra can be measured [47]. The kinetic energies of the photoelectrons in a UPS experiment typically range from about 15 to ˚ in this energy range, this 100 eV. Since the mean free path of electrons in matter is ∼10 A is an extremely surface sensitive technique. Therefore, the electronic and magnetic properties measured with spin-resolved UPS are from the outermost atomic layers of the crystal, which may differ from the bulk properties. Since CrO2 can reduce to Cr2 O3 under vacuum conditions at elevated temperatures, it is also important to monitor the structure of the surface with low energy electron diffraction (LEED) or scanning tunnelling microscopy (STM) to ensure that the surface has not converted to Cr2 O3 . Another factor that must be considered with UPS 6

J. Phys.: Condens. Matter 19 (2007) 315207

C A Ventrice Jr et al

Intensity (arb. un.)

(a)

6 4 2 Energy below EF (eV)

Intensities (arb. units)

8

0

tsp = 335s

100

Intensity (arb. un.)

tsp = 0s Au CrO2 × 40

0.4

0=EF

-0.4

Spin-Polarization (%)

(b) tsp = 120s

80 60 40 20

Energy below EF (eV)

0 5

4

3

2

Energy below EF (eV)

1

5

4

3

2

1

Energy below EF (eV)

Figure 4. (a) Photoelectron spectra of polycrystalline CrO2 films measured at 300 K with hν = 21.2 eV for different sputtering times (tsp ) with 500 eV Ne ions. Upper inset shows the larger binding energy scale for a spectrum after 335 s of sputter cleaning. Lower inset shows comparison between polycrystalline Au foil and CrO2 spectra (expanded by a factor of 40) in the vicinity of the Fermi level. (b) Spin polarization of photoelectrons after 120 s of sputter cleaning. Used with permission from [49].

experiments is the instrumental energy resolution. The resolution will depend on the type of light source, the energy of the photons, the kinetic energy of the photoelectrons, and the type, size, and pass energy of the electron analyser. Although an instrumental resolution of