throbber
Home Search Collections
`
`Journals About Contact us My IOPscience
`
`The effect of spatial confinement on magnetism: films, stripes and dots of Fe on Cu(111)
`
`This content has been downloaded from IOPscience. Please scroll down to see the full text.
`
`2003 J. Phys.: Condens. Matter 15 R1
`
`(http://iopscience.iop.org/0953-8984/15/2/201)
`
`View the table of contents for this issue, or go to the journal homepage for more
`
`Download details:
`
`IP Address: 128.237.144.75
`This content was downloaded on 21/12/2015 at 01:51
`
`Please note that terms and conditions apply.
`
`Lambeth Magnetic Structures, LLC Exhibit 2006
`
`LMBTH-000159
`
`

`
`INSTITUTE OF PHYSICS PUBLISHING
`
`JOURNAL OF PHYSICS: CONDENSED MATTER
`
`J. Phys.: Condens. Matter 15 (2003) R1–R30
`
`PII: S0953-8984(03)33359-4
`
`TOPICAL REVIEW
`
`The effect of spatial confinement on magnetism: films,
`stripes and dots of Fe on Cu(111)
`
`J Shen1, J P Pierce1,2, E W Plummer1,2 and J Kirschner3
`
`1 Solid State Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
`2 Department of Physics, University of Tennessee, Knoxville, TN 37996, USA
`3 Max-Planck-Institut f. Mikrostrukturphysik, Weinberg 2, D-06120 Halle/Saale, Germany
`
`Received 14 August 2002
`Published 20 December 2002
`Online at stacks.iop.org/JPhysCM/15/R1
`
`Abstract
`In this article, we review recent progress in the exploration of the complex
`magnetic phases of the fcc Fe/Cu(111) system. In particular, we emphasize
`the magnetic properties realized by the synthesis of novel nanostructures of
`Fe on Cu(111). These include monolayer films, one-dimensional stripe arrays
`and nanodot arrays. The effects of spatial confinement, together with strong
`spin–lattice correlations, result in dramatically different magnetic behaviour for
`the various manifestations of the Fe/Cu(111) system. Multi-scale theoretical
`calculations have been used to provide an understanding of the magnetic
`behaviour in each case.
`
`(Some figures in this article are in colour only in the electronic version)
`
`Contents
`
`Introduction
`1.
`2. Thermal MBE growth of Fe on Cu(111): rough and discontinuous films
`2.1. Morphology and structure
`2.2. Magnetism
`3. Laser MBE growth: two-dimensional ultrathin films
`3.1. Laser MBE growth
`3.2. Correlation between structure and magnetism
`4. Step decoration on vicinal surface: quasi-one-dimensional stripe array
`4.1. Formation of one-dimensional nanostructures
`4.2. Quasi-one-dimensional magnetism
`5. Buffer layer assisted growth: Fe dot assemblies on Cu(111)
`5.1. Formation of quasi-zero-dimensional clusters of Fe on Cu(111)
`5.2. Magnetic behaviour of arrayed dots
`6. Effect of spatial confinement on magnetism: direct comparison of films, stripes and
`dots of Fe on Cu(111)
`
`0953-8984/03/020001+30$30.00 © 2003 IOP Publishing Ltd
`
`Printed in the UK
`
`2
`3
`3
`6
`8
`8
`10
`13
`13
`15
`19
`19
`21
`
`24
`
`R1
`
`LMBTH-000160
`
`

`
`R2
`
`7. Summary and outlook
`Acknowledgments
`References
`
`1. Introduction
`
`Topical Review
`
`27
`27
`27
`
`In the last decade, basic science has had a remarkable impact on practical devices in the area
`of magnetic recording. In less than ten years, the discovery of giant magnetoresistance, a
`phenomenon that occurs in thin film sandwiches of magnetic and non-magnetic materials,
`developed into a $100 billion per year market for a new generation of hard disk drives.
`This amazingly rapid progress resulted from basic research that gave the community both
`the ability to grow these artificial structures and a basic understanding of their behaviour
`that allowed optimal tuning of their properties. The rich physics associated with these
`magnetic films [1, 2]—which are nanoscale in only one spatial dimension—provides ample
`testament that nanophase magnetic materials are not just smaller but also different! As efforts
`to reduce device size have continued, it has become imperative to investigate the magnetic
`properties of artificial structures on smaller length scales and in reduced dimensionality, as
`in nanowires [3–8], dots [9–12] and pillars [13]. These advances, coupled with emerging
`techniques for synthesizing magnetic nanowires [8, 14, 15], nanoparticles [16] and molecular
`magnets [17, 18], have established the research of spatially confined magnetic materials as a
`new frontier both in basic science and technology.
`New properties that emerge at the nanoscale have at least four origins:
`(1) As the surface-to-volume ratio increases, material properties are increasingly dominated
`by surface and interface effects—a 5 nm cube of bcc Fe contains ∼12 000 atoms, ∼2000
`of which are on the surface.
`(2) Spatial confinement results in new quantum phenomena. The oscillatory exchange
`coupling [19], GMR [20], spin-dependent tunnelling [21] and exchange bias [22–24]
`manifest in magnetic multilayers are linked with one or both of these factors.
`(3) A contribution that can perhaps be called ‘characteristic length effects’. The exotic effects
`seen in GMR spin valves, for example, would not be observed if the individual layers in
`the structure were thicker than the spin-diffusion length, which is the average distance that
`an electron will travel in a material before undergoing a scattering process that changes
`its spin.
`(4) In many spin systems, like colossal magnetoresistance (CMR) materials [25] and magnetic
`semiconductors [26, 27], correlation effects are already important in the bulk spin structure,
`spin fluctuations and spin transport.
`In general, spatial confinement will significantly
`change these correlation effects.
`
`Before the exotic magnetic and electronic properties of various nanostructures like
`ultrathin films, nanowires and nanodots can be explored, important strides must be made in
`controlling their synthesis. While state-of-the-art e-beam lithography may produce structures
`down to the nanometre scale, mass production of such structures by lithography or etching-
`based fabrication has proven to be exceptionally challenging [28, 29]. For this reason, in
`the past decade considerable effort has been devoted to investigating growth methods using
`self-assembly principles. Self-assembled magnetic nanowire and nanodot arrays have been
`achieved on various types of substrates.
`In this paper we will review studies of the spatial confinement effect on magnetism
`in a highly interesting system, Fe on Cu(111). We pick this particular system because it
`provides a classic demonstration of the tremendous impact that novel synthesis techniques
`
`LMBTH-000161
`
`

`
`Topical Review
`
`R3
`
`Figure 1. Schematics showing the phenomenon of twinning. The spacing between iron atoms on
`the Cu(111) surface is such that iron atoms in a particular island will either occupy ‘A’ sites only
`or ‘B’ sites only, depending on which of these two sites was settled upon by the atom that seeded
`the island. The image shows the fault line that forms when an A-type island tries to merge with a
`B-type island.
`
`can have on the study of nanomagnetism. Ultrathin films, nanowires and nanodots of Fe have
`been successfully grown on the Cu(111) substrate, despite the fact that conventional growth
`techniques like thermal evaporation and sputter deposition do not support the growth of any
`of these nanostructures. The fact that these three types of nanostructures, which are spatially
`confined in different dimensions,can be grown on a common template allows direct observation
`of the effect of spatial confinement on magnetism.
`Another major attraction of the Fe/Cu(111) system is that iron initially grows on this
`surface in the face-centred cubic structure, which is well known for its rich magnetic phases
`and strong spin–lattice coupling [30].
`In general, a small variation of the lattice constant
`or lattice distortion can result in drastic changes of magnetic phases that range from low-
`moment ferromagnetic phase, antiferromagnetic phase, ferrimagnetic phase and high-moment
`ferromagnetic phases. For the very same reason, ultrathin films of fcc Fe on the Cu(100)
`surface have been extensively studied in the last ten years, and many exciting magnetic phase
`transitions [31, 32] and strong spin–lattice correlations [33, 34] have been identified. The fcc
`Fe/Cu(111) system, as we will soon show below, is just as exciting and its degree of complexity
`is just as high. Combining the results for both crystallographic orientations of fcc Fe will also
`lead to a better overall understanding of the material.
`In order to emphasize the critical role of novel synthesis techniques in the study of low-
`dimensional magnetism, we organize this review as follows. In section 2 we will point out that
`conventional molecular beam epitaxy (thermal MBE) fails to yield any of the low-dimensional
`magnetic nanostructures. The structure and magnetism of the thermal MBE grown Fe/Cu(111)
`will be, nevertheless, discussed in that section to provide the readers with a reference point.
`In the following sections we discuss the growth, structure and magnetism of well ordered
`Fe/Cu(111) nanostructures. Ultrathin films, stripe arrays and dot assemblies of Fe on Cu(111)
`are described in sections 3–5, respectively. In section 6, we make a direct comparison of the
`magnetic properties of these manifestations of Fe on Cu(111). The final section provides a
`summary and an outlook on future research.
`
`2. Thermal MBE growth of Fe on Cu(111): rough and discontinuous films
`
`2.1. Morphology and structure
`
`When prepared with a conventional growth technique like thermal evaporation in ultrahigh
`vacuum, Fe has a strong tendency to form multilayer islands on Cu(111) due, in part, to an
`effect called twinning. Twin structures are common features of epitaxial growth on fcc(111)
`
`LMBTH-000162
`
`

`
`R4
`
`Topical Review
`
`50 nm
`
`5th ML
`4th ML
`3rd ML
`2nd ML
`
`0
`
`10
`
`20
`
`30
`
`substrate
`40
`50
`nm
`
`60
`
`70
`
`1.0
`
`0.8
`
`0.6
`
`0.4
`
`0.2
`
`0.0
`
`nm
`
`Figure 2. STM image of the rough morphology that results when two atomic layers of Fe are
`deposited via thermal evaporation on the Cu(111) surface at 220 K. The line scan below the image
`shows that this morphology is far from that of an ideal 2.0 ML film. The white arrows indicate the
`locations of ridge-like bcc(110) structures.
`
`surfaces [35]. As shown in figure 1, the rhombic surface unit cell of the fcc(111) surface
`provides two possible sites, A and B, for adatom nucleation. Since the difference between
`the nucleation energies of the two sites is generally very small, an adatom has essentially no
`preference for one site over the other. Those that nucleate at A sites seed the growth of A-type
`islands, and those that nucleate at B sites lead to B-type islands. These two types of islands
`cannot merge to form a smooth film because a fault line, as shown in figure 1, always exists at
`the boundary between them.
`A second feature that leads to roughness is low interlayer mass transport [36]. Because
`there is a high energy barrier for Fe atoms to overcome when moving from one atomic
`layer to the next, the thermal motion of the atoms is unable to ‘heal’ pits and peaks in the
`morphology. The consequences of fcc twinning and low interlayer mass transport can be seen
`in the scanning tunnelling microscope (STM) image of an Fe/Cu(111) film shown in figure 2.
`The film was prepared by MBE at a substrate temperature of 220 K, with the nominal Fe
`dosage of 2 monolayers (ML). The marked line profile shows that the typical island height is
`about 5 ML, and that a considerable fraction of the copper surface (darkest contrast) remains
`uncovered after two atomic layers of Fe are deposited.
`In figure 2, another noticeable feature is the appearance of some ridge-like islands (marked
`by white arrows). These ridge-like structures represent the typical morphology of bcc(110)
`structures, which are elongated due to one-dimensional lattice matching with the fcc Cu(111)
`
`LMBTH-000163
`
`

`
`Topical Review
`
`R5
`
`2.7 ML
`
`90 eV
`
`Figure 3. Ball model of Kujdimov–Sachs (KS) orientation. There are six Fe bcc(110) domains
`that match the Cu(111) substrate in the three (cid:2)110(cid:3) directions. At the lower left is the low-energy
`electron diffraction (LEED) pattern generated by a 2.7 ML thermal MBE-grown Fe on Cu(111)
`with KS orientation. The schematic at the lower right shows the satellite spots that form due to KS
`orientation.
`
`substrate. A ball model is shown in figure 3 to demonstrate lattice matching between bcc
`Fe(110) and fcc Cu(111). There are, in total, six possible domain configurations (Kurdjumov–
`Sachs or KS orientation) along the three (cid:2)011(cid:3) directions of Cu(111) surface. The six KS
`domains give rise to six satellite spots in the corresponding LEED patterns, as shown in the
`bottom pictures of figure 3 [37]. The morphological and structural evolution, as a function
`of thickness, is shown in the STM images and LEED data in figures 4 and 5, respectively.
`The data in these figures were acquired from Fe films grown on a Cu(111) substrate with a
`. The critical thickness of the fcc → bcc transition in these Fe
`◦
`slight miscut angle of 1.2
`films is between 2.3 and 2.7 ML, as evidenced by the appearance of bcc ridges in figure 4 and
`the change of interlayer distance (as determined by LEED-IV) in figure 5. In the past, there
`has been some inconsistency in values reported for the critical thickness of the fcc → bcc
`transition. A thickness of 5 ML was reported for Fe growth on a flat Cu(111) substrate [37],
`◦
`while 1.5 ML was found for Fe on a vicinal Cu(111) substrate with an 8
`miscut angle [38].
`This inconsistency, however, reflects the fact that the nominal thickness is only an average
`value of the height of an assembly of multilayer islands. STM studies indicate that, when the
`local thickness of a particular island exceeds 6 ML, the morphology of that island becomes bcc-
`
`LMBTH-000164
`
`

`
`R6
`
`Topical Review
`
`1.8 ML
`
`2.3 ML
`
`60 nm
`
`60 nm
`
`2.7 ML
`
`5.4 ML
`
`60 nm
`
`60 nm
`
`Figure 4. STM images of thermal MBE-grown Fe on Cu(111) that show the morphological change
`that occurs when the film undergoes an fcc to bcc structural transition between nominal thicknesses
`of 2.3 and 2.7 ML.
`
`like with elongated structures. As we will discuss later, Fe adatoms have a strong preference
`for nucleation at the atomic step edges of a Cu(111) substrate. This implies that the height
`distribution of the mutilayer islands is strongly affected by the density of the substrate steps,
`even if the nominal thickness of Fe is unchanged. With this in mind, most of the varying results
`in the literature are actually consistent with each other.
`The substrate temperature has a profound influence on the growth and especially the
`fcc → bcc transition of Fe on Cu(111). A forward-scattering x-ray photoelectron diffraction
`study showed that the fcc structure of Fe on Cu(111) can only be obtained when the substrate is
`kept warmer than 80 K during growth [39]. At temperatures lower than 80 K, epitaxy of Fe on
`Cu(111) cannot be properly established and the more stable bcc phase dominates the growth.
`At temperatures significantly above room temperature, ∼700 K for example, the bcc Fe phase
`only appears at a thickness greater than 40 ML. However, the seemingly delayed fcc → bcc
`phase transition is a direct consequence of Fe–Cu interdiffusion. In fact, Fe–Cu interdiffusion
`already occurs to some extent at room temperature [40].
`
`2.2. Magnetism
`
`The thermal MBE grown Fe/Cu(111) films have consistently been found to be in a low-
`moment phase, in stark contrast to the high-moment phase observed in thermal MBE grown
`Fe films on the Cu(100) surface [41]. Early experiments on surface-coated Fe/Cu(111) films,
`−8 Torr vacuum, showed that the magnetic moment of the fcc Fe is only about
`prepared in a 10
`0.58 µB [42]. Electron-capture spectroscopy measurements on UHV-grown, uncoated films
`detected short-range ferromagnetic order in films thinner than 2 ML [43], which is consistent
`with the multilayer island morphology and the discontinuity of the films shown in figure 2.
`More recently, the thickness dependence of magnetism of Fe/Cu(111), measured by surface
`
`LMBTH-000165
`
`

`
`Topical Review
`
`R7
`
`0.21
`
`0.208
`
`3.0 ML
`
`0.206
`
`"fcc"
`
`"bcc"
`
`0.204
`
`0.202
`
`1.5 ML
`
`interlayer distance (nm)
`
`0.2
`
`0
`
`2
`
`6
`4
`thickness (ML)
`
`8
`
`10
`
`Figure 5. Plot of the vertical lattice constant of thermal MBE-grown Fe films as a function of
`thickness. A sharp and sudden drop in this lattice constant is observed at a thickness of 2.5 atomic
`layers, which corresponds well to the thickness at which the ridge-like structures appear in the
`STM images in figure 4.
`
`Fe on Cu(111)
`
`fcc
`
`bcc
`
`saturation magnetization (a.u.)
`
`0
`
`coverage (ML)
`
`Figure 6. Thickness dependence of the surface SMOKE signal observed when the magnetization
`of thermal MBE-grown Fe films is saturated in a direction perpendicular to the Cu(111) surface.
`The data show that a low-moment to high-moment magnetic transition occurs at the same thickness
`as the structural changes discussed in figures 4 and 5.
`
`magneto-optical Kerr effect (SMOKE), gave strong evidence that the uncoated fcc Fe has a
`considerably smaller net magnetic moment than that of bulk bcc Fe [44]. Figure 6 shows the
`saturation polar Kerr intensity as a function of Fe/Cu(111) thickness. Apparently, the positive
`slope of Kerr intensity as a function of thickness takes two distinctly different values in the fcc
`
`LMBTH-000166
`
`

`
`R8
`
`Topical Review
`
`Figure 7. Hysteresis loops obtained with the SMOKE for thermal MBE-grown Fe films of various
`thicknesses. The data indicate that the easy axis of magnetization of these films rotates from the
`perpendicular to the in-plane direction at the thickness at which the previously discussed structural
`transition happens. The vertical scale of the plot of the in-plane loops has been reduced relative
`to that of the plot on the left since the intensity observed in the longitudinal Kerr effect is smaller
`than what is seen in the polar geometry.
`
`and bcc regime, with the bcc slope being nearly four times larger. Assuming that the saturation
`Kerr intensity is roughly proportional to the net magnetic moment of the sample, one would
`conclude that the MBE-grown fcc Fe/Cu(111) has a moment of the order of 0.5 µB. A number
`of photoemission studies show signs of smaller exchange splitting for the films in the fcc
`regime than that of the films in the bcc regime [45–47]. This is consistent with the observation
`of a low-moment phase. Recently, further details regarding the nature of this low-moment
`phase have been revealed by an x-ray magnetic circular dichroism (XMCD) study [48], which
`will be discussed in section 4.
`The fcc → bcc structural transition not only induces the low-moment to high-moment
`magnetic transition, but also drives a spin reorientation transition. Figure 7 shows the side-
`by-side comparison of polar and longitudinal MOKE hysteresis loops of the same Fe/Cu(111)
`sample measured in figures 4 and 5. At thicknesses below 2.3 ML, only polar loops can be
`measured. The polar loops generally have a very low remanence,reflecting the reduced thermal
`stability of the multilayer islands in the system. When the thickness is higher than 2.3 ML,
`in-plane hysteresis loops become detectable, while the saturation field for polar loops increases
`rapidly with increasing thickness. Well-defined square hysteresis loops are measured in the
`in-plane direction when the films are in a bcc structure.
`
`3. Laser MBE growth: two-dimensional ultrathin films
`
`3.1. Laser MBE growth
`
`While Fe/Cu(111) films grown with conventional molecular beam epitaxy show interesting
`features such as a low-moment fcc Fe phase and a perpendicular to in-plane spin reorientation,
`the multilayer island morphology does not have a well-defined dimensionality and is thus not
`
`LMBTH-000167
`
`

`
`Topical Review
`
`R9
`
`Figure 8. Schematic showing how laser MBE can be incorporated into a surface-analysis system.
`The laser is directed through a window and into the vacuum chamber where it ablates a target
`material onto a substrate that is held nearby.
`
`ideal for studying the effect of spatial confinement on magnetism. In the last decade, a number
`of techniques have been developed for inducing the self-assembly of high quality magnetic
`nanostructures on a given template. In this section we will discuss a very powerful method for
`growing two-dimensional ultrathin films: laser molecular beam epitaxy.
`Laser MBE essentially incorporates most of the growth principles of traditional pulsed
`laser deposition (PLD). In PLD, a powerful, nanosecond-pulsed excimer laser is focused onto
`a target. After a complex laser–solid interaction [49], a plasma plume is generated from the
`target and expands quickly toward the substrate. The plasma plume consists mainly of neutral
`atoms from the target that have a relatively moderate energy (∼1 eV), plus a small fraction of
`ions whose energy can be as high as ∼100 eV. A remarkable feature of PLD is that it yields a
`−1 or higher, which is nearly
`deposition rate (during each laser pulse) of the order of 106 ML min
`6 orders of magnitude higher than that of thermal MBE. This high deposition rate, according
`to growth theory [50, 51], tends to enhance the nucleation density, which in turn favours two-
`dimensional growth. This mechanism is basically similar to the so-called reentrant growth at
`low substrate temperatures, which also leads to high nucleation densities. Attempts to produce
`smooth Fe films using conventional MBE at low substrate temperatures, however, fail since Fe
`forms in the bcc structure from the initial stages of growth [39]. Once the bcc structure forms,
`Fe films will not grow two-dimensionally, or in a layer-by-layer mode, because of the large
`lattice mismatch between bcc Fe (of any surface orientation) and fcc Cu(111). The typical
`rough morphology of the bcc Fe on Cu(111) has already been shown in figure 5.
`The features that distinguish a laser MBE system from a PLD system include the ultrahigh
`vacuum environment and the use of reflection high-energy electron diffraction (RHEED) to
`monitor the growth of individual atomic layers in real time. Figure 8 shows a schematic of
`a typical laser MBE system that has been integrated into a surface analysis chamber at Oak
`Ridge National Laboratory that is equipped with a number of other state-of-the-art tools for
`in situ characterization of film morphology, structure and magnetism. The combination of
`layer-by-layer RHEED intensity oscillations and STM analysis of layer fillings is by far the
`most accurate method for thickness calibration. It was shown recently that such a reliable
`thickness calibration method played a critical role in the determination of the hotly debated
`antiferromagnetic spin structure of the fcc Fe/Cu(100) system [52].
`With the help of the laser MBE growth, nearly ideal layer-by-layer growth of fcc Fe
`films can be achieved, which is in stark contrast to the multilayer island morphology of the
`Fe/Cu(111) produced by thermal MBE. The surface morphology of a single atomic layer of
`
`LMBTH-000168
`
`

`
`R10
`
`Topical Review
`
`Figure 9. STM images that allow us to compare the morphologies of a single atomic layer of Fe
`as grown on Cu(111) with thermal MBE (left) and laser MBE (right). The laser MBE-grown film
`is a nearly perfect atomic layer. A substrate step runs from top to bottom through the centre of the
`image on the right.
`
`Fe/Cu(111) grown by laser MBE and thermal MBE is compared in figure 9. Both films were
`prepared at an identical substrate temperature of 220 K. The thermal MBE film, as discussed
`earlier, exhibits typical multilayer island morphology that leaves a considerable amount of Cu
`substrate uncovered. The laser MBE Fe film, at nominal thickness of 1 ML, covers about
`95% of the substrate, which can be considered as a nearly perfect monolayer film. This two-
`dimensional layer-by-layer growth persists up to a film thickness of 6 ML, above which an
`fcc → bcc structural transition occurs [53]. As we have discussed earlier, for the thermal
`MBE-grown Fe/Cu(111), the bcc structure starts to form locally in multilayer islands that are
`6 ML high. The consistency of the fcc → bcc critical thickness for the thermally deposited
`and laser MBE-grown Fe/Cu(111) films implies that 6 ML and below is the thickness range in
`which the fcc phase of Fe is stable on a Cu(111) substrate.
`
`3.2. Correlation between structure and magnetism
`
`Laser MBE-grown Fe/Cu(111) films exhibit multiple magnetic phase transitions. Unlike
`thermal MBE-grown Fe/Cu(111), the laser deposited films appear to have a high-moment
`phase when the thickness is less than 3 ML. Above 3 ML, this high-moment phase transforms
`into a low-moment phase, and eventually becomes a high-moment phase again once the films
`complete the fcc → bcc structural transition (>6 ML). Figure 10 shows the magnetic phase
`diagram for laser MBE-grown Fe/Cu(111) films. The change in saturated polar Kerr intensity
`with thickness for films that are thinner than 3 atomic layers is more than 3 times higher than
`that for films thicker than 3 ML. At thicknesses beyond those shown in the plot, when the
`films become bcc (above 6 ML), the slope of the Kerr intensity almost recovers the original
`value (not shown here). Again, assuming the Kerr intensity is proportional to the net moment
`of the films, one would conclude that at low thicknesses (<3 ML), the laser MBE-grown fcc
`Fe/Cu(111) films have a high magnetic moment which more or less equals the bcc Fe moment
`(2.2 µB). This high-moment phase transforms into a low-moment phase of 0.7 µB when the
`film thickness exceeds 3 ML. To date, the nature of the low-moment phase, i.e. whether it
`comes from a low-spin ferromagnetic phase, a ferrimagnetic phase or even a noncolinear spin
`structure [54], is undetermined.
`
`LMBTH-000169
`
`

`
`Topical Review
`
`R11
`
`1
`
`4
`3
`2
`thickness (ML)
`
`250
`
`200
`
`150
`
`100
`
`50
`
`0
`
`Polar Kerr intensity
`
`Curie temperature (K)
`
`Figure 10. The upper plot shows the magnetization of the laser MBE-grown films as a function
`of thickness. The spin reorientation that occurs at a film thickness of 2 ML does not appear to be
`associated with the high-moment to low-moment transition that is observed at a thickness of 3 ML.
`The lower plot shows the thickness dependence of the Curie temperature of these films.
`IV/LEED
`RHEED
`
`TD
`
`PLD
`
`bcc Fe
`
`fcc Cu
`
`2,30
`
`2,25
`
`2,20
`
`interrow distance (Å)
`
`0
`
`2
`
`4
`6
`8
`thickness (ML)
`
`10
`
`12
`
`fcc
`
`bcc
`
`0
`
`2
`
`10
`8
`6
`4
`Thickness (ML)
`
`12
`
`14
`
`2.11
`
`2.10
`
`2.09
`
`2.08
`
`2.07
`
`2.06
`
`2.05
`
`2.04
`
`2.03
`
`Interlayer distance (A)
`
`Figure 11. Structural measurements of laser-deposited Fe films that show a sharp decrease in
`the vertical lattice constant (left) and a sudden increase in the in-plane lattice constant (right) at a
`thickness of 6.0 atomic layers. The decrease of the in-plane lattice constant between 4 and 6 ML
`may be associated with the high-moment to low-moment transition, as discussed in the text.
`
`LMBTH-000170
`
`

`
`R12
`
`Topical Review
`
`band energy
`dipolar energy
`total energy
`
`1
`
`4
`3
`2
`thickness (ML)
`
`5
`
`6
`
`1
`
`0.5
`
`0
`
`-0.5
`
`Magnetic anisotropy energy (meV)
`
`-1
`
`0
`
`Figure 12. Plot of the results of an ab initio calculation of the various contributions to the magnetic
`anisotropy of laser MBE-grown Fe films on Cu(111). The change in sign of the total anisotropy
`energy that is found at 2.5 ML indicates that a perpendicular to in-plane spin flop should occur at
`that thickness. This is in good agreement with experimental observations (courtesy of B Ujfulussy).
`
`Detailed structural analysis indicates that the high-moment to low-moment phase transition
`is associated with a slight change of lattice constant of the Fe films. Figure 11 shows the
`thickness-dependent vertical (a) and lateral (b) lattice constants of the laser MBE-grown
`Fe/Cu(111) films. The vertical lattice constant, or the interlayer distance, was determined
`by LEED-IV measurements. It does not change throughout the entire fcc thickness regime.
`The lateral lattice constant, or the inter-row distance, was determined by RHEED. It initially
`increases with thickness, reaching a maximum between 3 and 4 atomic layers. The distance
`between rows then decreases with increasing thickness, before it adopts a much larger value in
`the bcc regime (>6 ML). Figure 11(b) clearly shows that the thinner Fe films (<4 ML) have
`a larger inter-row distance than that of the thicker Fe films (between 4 and 6 ML). The initial
`increase of the inter-row distance likely reflects the fact that some of the electrons that contribute
`to the diffraction pattern actually diffract from the Cu substrate itself. This effect would make
`the measured value of the Fe lattice constant appear smaller during the initial stages of growth,
`and becomes less important with increasing thickness, since the penetration depth of a RHEED
`beam is typically no more than 2–3 atomic layers. The average inter-row distance peaks at
`3–4 ML, at which the contribution from the substrate is negligible. The subsequent decrease
`of the inter-row distance, however, has to be attributed to the intrinsic structural behaviour of
`the Fe films, since the presence of the substrate no longer affects the measured values. With
`this argument, and the information from both figures 11(a) and (b), one may conclude that, in
`the fcc thickness regime, the atomic volume of Fe films is slightly larger at low thicknesses
`(<3–4 ML) than it is when the films are between 4 and 6 ML thick. Considering the well-
`known fact that a larger lattice constant favours a larger magnetic moment for fcc Fe [30], the
`lattice constant decrease between 4 and 6 ML in figure 11(b) provides a plausible explanation
`of the high-moment to low-moment phase transition for the laser MBE-grown films.
`A perpendicular to in-plane spin reorientation transition was also observed in the laser
`MBE-deposited Fe/Cu(111) films. This occurs at about 2 ML and thus is not associated with
`
`LMBTH-000171
`
`

`
`Topical Review
`
`R13
`
`either the fcc → bcc structural transition (6 ML) or the high-moment to low-moment magnetic
`transition (3 ML) in this system. This is in stark contrast to the spin flop transition observed
`in thermal MBE-grown Fe/Cu(111), which is originated by an fcc → bcc structural phase
`transition. An ab initio calculation, based on the fully relativistic screened KKR method, was
`recently made by Uljfulussy et al [55], showing that this spin flop transition is governed by the
`balance between perpendicular magnetocrystalline anisotropy and in-plane shape anisotropy.
`As shown in figure 12, the magnetic anisotropy band energy (magnetocrystalline anisotropy)
`has positive values that favour perpendicular magnetization. The weak thickness dependence
`implies that the main contributions to this energy term come from the surface and interface.
`The demagnetizing energy (shape anisotropy), as is the case in all magnetic ultrathin films,
`has negative values and increases linearly with thickness. The total anisotropy energy, i.e. the
`sum of the band energy and demagnetizing energy, changes its sign from positive to negative
`at a thickness of about 2.5 ML, predicting a magnetization reorientation at that thickness. This
`is in good agreement with the experimental findings.
`The unusual thickness dependence of the Curie temperature (TC ) of the Fe ultrathin films,
`which seems to have a maximum at about 2 ML (figure 10), is a puzzle that is still not
`understood. Although some kind of non-monotonic behaviour of TC is expected due to the
`magnetic phase transitions, there are two major surprises regarding its thickness dependence
`in this system. First, TC starts to decrease at 2 ML, a thickness at which the Fe films are
`still uniformly in a high-moment phase. Second, TC monotonically decreases with increasing
`thickness above 2 ML. At that thickness, this behaviour cannot be attributed to the onset
`of the low-moment phase or to the spin reorientation. Interestingly, the first behaviour was
`also observed in the Fe/Cu(100) system, and was understood to be caused by a temperature-
`induced structural transition during the TC measurement [56]. Temperature-induced structural
`transitions have not been carefully examined so far in the Fe/Cu(111) system and cannot be
`used to explain the totally unexpected, monotonic decrease of TC. Further structural and spin-
`polarized electronic information is needed to understand the abnormal behaviour of TC in this
`system.
`
`4. Step decoration on vicinal surface: quasi-one-dimensional stripe array
`
`4.1. Formation of one-dimensional nanostructures
`
`One-dimensional (1D) systems have intrigued scientists for many years. In the years before
`powerful and cheap computing, 1D models, where analytical solutions existed, often appeared
`to be the best theoretical vehicle for studying many phenomena. Physics in one dimension
`is quite different from physics in two or three. For example, in 1D, there can be no phase
`transitions or long-range order [57] nor can there be electrical conduction because all electronic
`states are localized [58]. We live, however, in a space of three dimensions, so

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket