"fa-$2? as? 23M “32:94.1 ragga» 33: V‘Tfiafii ' W'h r‘v’“ '13 '4 k-A ' -.-.-. . ."”-'. ‘I‘ L- n q» ,.....-»‘ v n .. .-. (. mun" HM-.«-- (.2,ng I .Ezgmki N. 7.‘ .2 I)- (“I ———————__..... — o‘- v-—. r. LI mi I: Y I’licrtgaa Stye I ".r,t ; ‘ -""A- ‘4. 0 53“." “1"". I uni. \t..~l‘.,1g‘pl J This is to certify that the dissertation entitled HRTEM AND EELS STUDIES OF NANOSCALE STRUCTURED ELECTRONIC MATERIALS presented by JIAMING ZHANG has been accepted towards fulfillment of the requirements for the Materials Science and Engineering Ph.D. degree in flZ/Kfl Major Professor’s Signature ,7//7/7 7 Date MSU is an affirmative-action. equal-opportunity employer ._—_—-.--—.- PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DAIEDUE DATEDUE DAIEDUE AUHI % 01220433 6/07 p:lClRC/Date0ue.indd—p.1 HRTEM AND EELS STUDIES OF NANOSCALE STRUCTURED ELECTRONIC MATERIALS By Jiaming Zhang A DISSERTATION Submitted to Michigan State University in partial fulfilment of the requirements for the degree of DOCTOR OF PHILOSOPHY Materials Science and Engineering 2007 ABSTRACT HRTEM AND EELS STUDIES OF NANOSCALE STRUCTURED ELECTRONIC MATERIALS By Jiaming Zhang This thesis focuses on the growth and structure of a number of nanoscale structured electronic materials characterized using HRTEM and EELS. Both rare earth silicide nanostructures self-assembled on Si(001) and sputtered GMR multilayers have been studied by characterizing their crystal structures, interfacial epitaxy, interfacial chemical nature, and electronic nature, which provide fundamental insights into material behavior at the nanometer scale. High aspect ratio nanowires and nanosized islands have been observed to self- assemble on Si(OOl) for both the Gd and Tm silicide systems. HRTEM results Show that Gd silicide nanostructures exhibit either the hexagonal GdSiz.x or the orthorhombic GdSiz crystal structures, with lattice parameters consistent with the bulk phases. In the case of Tm, the observed nanostructures are likewise either hexagonal or orthorhombic. The hexagonal phase has lattice parameters consistent with the bulk, while the orthorhombic does not. For both systems, bi-phasic silicide structures were observed, which may reflect a mechanism for strain accommodation at the interface with the substrate. In the case of Gd, the phase with lower strain lies at the substrate. For the case of Tm, the relative mismatches of the two phases predicted from bulk silicide lattice parameters disagree with that derived fi'om measured lattice constants, and it is a relaxed orthorhombic phase at the interface that appears to have the lowest mismatch with the substrate. EELS studies were carried out to compare the electronic structures of metallic Gd, thin film Gd silicide, and Gd oxide in bulk phases and Gd silicide nanostructures. The results from the three bulk phases are similar, while the intensity ratio of M52M4 in the GdSiz nanostructures varies from the bulk, which may suggest that a slightly different spin state exists in the silicide nanostructures. As-sputtered and annealed F(Co, or Py)/Al multilayers have been studied using HRTEM and EELS. Although the interfacial intensity profiles from EELS spectrum images suggest some limited intermixing exists in the F/Al interfacial regions, both HRTEM and diffraction studies Show no obvious intermediate phase formation. In particular, the annealing treatments do not significantly alter the multilayer structures. In contrast, intermediate phase formation has been observed in both Cu/Al multilayers and Cu/Al regions in spin valves. Tetragonal AlgCu and bcc AlCu3 are formed in the Cu(8nm)/Al(10nm) multilayers, while A12Cu and fee Cu are formed in the Cu(5nm)/Al(3nm) multilayers. For Cu/Al/Cu layers in the spin-valves, evidence of AlzCu and AlCu3 phase formation in the annealed spin-valve with the 30m Al layer was found, while A12Cu and Cu were observed in the as-sputtered spin-valve with the lOnm Al layer. These results are discussed in terms of the balance between interfacial and volume fi'ee energies in order to rationalize the formation of non-equilibrium structures. ACKNOWLEDGMENTS First of all, I wish to express my gratitude to my research advisors, Professor Martin A. Crimp and Professor Jun Nogami throughout my entire PhD. research. In 2003, I joined Prof. Crimp and Nogami’s team and started my research on rare-earth silicide nanowires under their supervision. They spent enormous time and energy to help me to start as a self-motivated graduate student. I am especially grateful for their patience and encouragement during my Ph.D. study. I feel very lucky to learn various advanced microscopy techniques from Prof. Crimp, which I believe will help me in my future career. I am also gratefiil for Prof. Nogami’s time and effort to come from Toronto to East Lansing several times after he moved to U of Toronto to help me with my research. Without Prof. Crimp and Prof. Nogami’s huge contribution, I do not think my dissertation is possible. I am grateful to my committee members, Prof. Case, Prof. Norman Birge and Prof. Tim Hogan for their valuable suggestions and comments on my research and dissertation. I also express my appreciation to Dr. Xudong Fan (Center for Advanced Microscopy) for his excellent TEM technical. I am also grateful to Prof. Bass and Prof. Pratt for providing me with GMR multilayer samples for my Ph.D. research. I feel very lucky that I have been able to study multiple systems during my study. I would also like to thank my previous and current co-workers Dr. Gangfeng Ye, Dr. Yucheng Lan, Yan Cui and Dr. Chigusa Ohbuchi. I enjoyed working with them, and their help, encouragement, and humor will remain in my memory. iv Finally, I would thank my family members, my wife Ran Mu, my parents and my in-laws. They have spent extensive time and effort to support me and take care of my newborn daughter Chelsea Zhang. Their immeasurable love and understanding are overwhelming and go beyond words. TABLE OF CONTENTS LIST OF TABLES ............................................................................... viii LIST OF FIGURES ................................................................................. x Chapter 1 Overview ................................................................................ 1 Chapter 2 Background ............................................................................. 3 2.1 Fundamental of HRTEM and EELS ...................................................... 3 2.1.1 HRTEM ...................................................... ' ....................... 4 2.1.1.1 Specimen Function ................................................... 4 2.1.1.2 Microscope Effect .................................................... 5 2.1.1.3 Fast Fourier Transform .............................................. 7 2.1.2 EELS .............................................................................. 8 2.1.2.1 Fundamentals ......................................................... 8 2.1.2.2 Energy-Loss Near-Edge Structures in EELSIO 2.2 Self-Assembled Silicide Nanowires .................................................... 12 2.2.] Introduction to Self-Assembled Silicide Nanowires ...................... 12 2.2.2 Crystallography of Epitaxial Silicide Nanowires .......................... 13 2.2.2.1 General Understanding of Bulk RE Silicides Phases ............ 13 2.2.2.2 Self-Assembled RE Silicide Nanostructures. .16 2.2.2.3 Crystal Structure of Bulk Phases in the Gd-Si and Tm-Si Systems ................................................................ 20 2.2.3 Electronic Structure of Silicide Nanowires..................................24 2.2.3.1 EELS Study of Electronic Structure .............................. 25 2.3 Giant Magnetoresistance (GMR) Multilayers .......................................... 27 2.3.1 Introduction of the GMR Effect .............................................. 27 2.3.1.1 Origin of the GRM Effects ......................................... 27 2.3.1.2 CIP and CPP ......................................................... 29 2.3.1.3 Exchanged-Biased Spin-valves .................................... 30 2.3.1.4 Current-Induced Magnetization Switching ...................... 30 2.3.2 Correlation between GMR and Multilayer Structures ..................... 31 2.3.3 Motivation of This Study ...................................................... 32 Chapter 3 Experimental Approaches and Procedures ................................... .39 3.1 Bulk Gd Oxide, Metallic and Silicide Reference Sample ............................. 39 3.1.1 Bulk Gd203 ....................................................................... 39 3.1.2 Bulk Metallic and Silicide Samples Produced by Dc-magnetron Sputtering ........................................................................ 40 3.2 RE Silicide Nanostructures Grown in UHV Chamber ................................. 40 3.3 Magnetic Multilayers Synthesis .......................................................... 42 3.4 TEM Characterization ..................................................................... 42 vi 3.4.1 Cross-sectional TEM Sample Preparation ................................... 43 3.4.2 High Resolution TEM45 3.5 EELS Characterization ..................................................................... 46 3.5.1 Design of Electron Energy-Loss Spectrometers ........................... 46 3.5.2 EELS Spectrum Collection ................................................... 47 3.5.3 Gd M-Edge Spectrum Curve Fitting ........................................ 48 3.5.4 Spectrum Image Collection .................................................... 48 3.5.5 Intensity Profiles and Simulation ............................................ 59 Chapter 4 Results from RE Silicide Nanostructure Studies ............................. 51 4.1 Gd Oxide, Bulk Metallic and Silicide Crystal Structure Study ..................... 51 4.2 HRTEM Study on Gd and Tm Silicide Nanostructures .............................. 56 4.2.1 Gd Silicide Nanostructures .................................................... 56 4.2.2 Tm Silicide Nanostructures ................................................... 64 4.2.2.1 Tm Silicide Nanostructure Morphologies ....................... 64 4.2.2.2 HRTEM Image Simulation ....................................... 66 4.2.2.3 Results from HRTEM Imaging ................................... 68 4.3 EELS Study on Gd Silicide Nanostructures ........................................... 76 4.4 Conclusions .................................................................................. 80 Chapter 5 Results from Multilayer Structure Studies82 5.1 Co(Py)/Al Multilayers .................................................................... 82 5.1.1 AS-sputtered and Annealed Co/Al Multilayers ............................. 82 5.1.2 As-sputtered and Annealed Py/Al Multilayers ............................. 95 5.2 Cu/Al Multilayers and Cu/Al in Spin-valve Structure .............................. 100 5.2.1 Cu/Al Multilayers ............................................................. 101 5.2.1.1 Cu(8nm)/Al(10nm) Multilayer Structures ...................... 101 5.2.1.2 Cu(3nm)/Al(5nm) Multilayer Structures ....................... 108 5.2.2 Cu/Al in Spin-Valve Structure ............................................... 110 5.2.2.1 As-Sputtered Spin-Valve with 10m Al ........................ 110 5.2.2.2 Annealed Spin-Valve with 30nm A1 ............................ 118 5.3 Discussion and Coclusions ............................................................... 125 Chapter 6 Overall Discussion and Summary ............................................... 128 Chapter 7 Conclusions and Suggestions for Future Works ............................. 136 7.1 Patterned Growth Studies of Silicide Nanostructures ................................ 136 7.2 High-resolution EELS Analysis from Monochromated TEM ....................... 137 7.3 EELS Quantification Analysis in Magnetic Multilayers ............................. 138 REFERENCES ................................................................................... 139 APPENDIX A Detailed TEM Sample Description ....................................... 147 vii Table 2.1 Table 2.2 Table 2.3 Table 2.4 Table 2.5 Table 2.6 Table 2.7 Table 2.8 Table 4.1 Table 4.2 Table 4.3 Table 5.1 Table 5.2 Table 5.3 Table 5.4 LIST OF TABLES Important edges in EELS associated with their electronic states excitation and degeneracy .................................................................... 10 A summary of the lattice constants of RE Silicides and their growth behavior on Si(001) ............................................................... 13 Gd-Si crystal structure data [19] ................................................. 15 Gd disilicide phases and lattice mismatches with Si<110> after [19].. . ...22 Tm silicide crystallographic data and lattice mismatches with Si<110> after [19] ............................................................................. 23 White lines energy differences and intensity ratios for the investigated rare earth oxides 26] .................................................................... 26 Co-Al crystal structure data [19] .................................................. 35 Cu-Al crystal structure data [19] ................................................. 38 Lattice parameters of the bulk GdSi2 phases and those measured from the three imaged nanostructures. (I, II, III refers to the imaged nanostructures) ...................................................................... 58 Lattice parameters of the bulk Tm silicide phases and those measured from the imaged nanostructures ........................................................ 71 Summary of N and M peak positions and intensity rati measurements of the white lines from the metallic Gd, the bulk GdSiz, the GdSiz nanostructure (NS), and Gd203 powder in comparison to Gd203 from EELS Atlas [27] ................................................................... 80 Lattice parameters and close-packed interplaner spacing variants for the multilayers ......................................................................... 88 Summary of the lattice spacings measured from all the phases in the as- sputtered spin-valve ............................................................. 104 Summary of the lattice spacings measured from all the phases in the as- sputtered Spin-valve ............................................................. 1 15 Summary of the lattice spacings measured from all the phases in figure 5.29 ................................................................................ 123 viii Table 6.1 Table 6.1 Electrical resistivity data of the phases in the Cu/Al multilayers, after [73] ........................................................................... 135 ix LIST OF FIGURES Chapter 2 Figure 2.1 (a) sin x(u) vs. u (without damping), and (b) T(u) vs. II modified by imposing an envelope function (dashed line); Af = -100nm, C5 = 2.2 mm ................................ 6 Figure 2.2 Figure 2.2 (a) An HRTEM image of Si along [110] zone (b) corresponding FFT pattern. The reciprocal spacings of the spots are related to the corresponding Si atomic planes .......................................................................................... 7 Figure 2.3 A typical EEL spectrum fiom an NiO specimen including zero-loss peak, plasmon peak, and compositional edges for NiO ................................................. 9 Figure 2.4 Gd-Si phase diagram, after [19] ...................................................... 14 Figure 2.5 Tm-Si phase diagram, after [19] ...................................................................... 16 Figure 2.6 Schematic representation of a hexagonal Silicides forming a nanowire on Si(001) due to the small mismatch along the a axis and large mismatch along the c axis, after [20] ..................................................................................................... 17 Figure 2.7 AFM image of GdSiz nanowires and nanoislands grown on Si(OOI) substrate along two perpendicular Si <110> directions, after [17] ....................................... 18 Figure 2.8 Plot of hexagonal Ale type of RE silicide anisotropic mismatches with Si(001). With respect to Si(001) two perpendicular <110> directions, the lattice mismatches of Silicides have a large absolute value between 4% and 10% and a small lattice mismatch within i2%. (after [11]) ....................................................... 19 Figure 2.9 An AF M image of Tm silicide nanostructures on Si(001) with 0.98 ML Tm coverage .............................................................................................. 20 Figure 2.10 Crystal models of RE Silicides, (a) hexagonal Ale structure (b) orthorhombic GdSiz or tetragonal ThSiz structure. (after [20] ) .............................. 21 Figure 2.11 Projections along various axes of the hexagonal and orthorhombic GdSiz phases ................................................................................................. 22 Figure 2.12 Projections along various axes of the hexagonal Tm3Si5 and orthorhombic TmSi phases ................................................................................................. 24 Figure 2.13 Magnetoresistance curves of Fe/Cr multilayers at 4.2 K [32] .................. 28 Figure 2.14 Schematic representations of the scattering electrons in a magnetic multilayer. The signs + and - are for spins Sz=+l/2 and Sz=-1/2, respectively. F represents a ferromagnetic layer and N represents a non-magnetic layer, after [33] ...................... 29 Figure 2.15 Schematic representative of a spin-valve structure ............................... 30 Figure 2.16 Co-Al phase diagram [19] ........................................................... 34 Figure 2.17 Cu-Al phase diagram [19] ........................................................... 37 Chapter 3 Figure 3.1 Schematic representation of RE Silicide nanostructure growth experimental setup ................................................................................................... 41 Figure 3.2 Schematic representation of the process of preparing a cross-sectional TEM sample ................................................................................................. 45 Figure 3.3 Schematic overview of the in-column Q filter (a) and post-column Gatan Image Filter (b) after [2] ........................................................................... 47 Figure 3.4 Zero-loss peak with a full width at half maximum (FWHM) ~1.0 eV shows that the EELS spectrometer is well aligned ...................................................... 48 Figure 3.5 A typical EELS spectrum of M45 edge (black line) fiom GdSiz fitted as a superposition of a power-law background (straight line), cross sections from Hartree- Slater model (rising line), and a set of Lorentzian profiles (gray lines). The dotted curve Shows a good curve fitting result ................................................................. 49 Figure 3.6 Simulation of an F(Co, Ni, or Fe) L-edge intensity profile with a 0.5nm FWHM of Gaussian electron beam profile across idealized F/Al interface ................ 50 Chapter 4 Figure 4.1 X—ray diffraction shows the bcc Gd203 agree with the prototype of Mn203 (Bixbyite) structure ................................................................................. 52 Figure 4.2 (a) Phase contrast TEM image of Gd203 powder particles on a holy carbon film. Insets (b) and (c) Show HRTEM images of a single grain and poly-grain atomic lattice fi'om two powders respectively .......................................................... 52 Figure 4.3 (a) Phase contrast TEM image and (b) corresponding selected area electron diffraction pattern of amorphous Si and metallic Gd layers alternately sputtered on a Si(001) substrate ................................................................................... 53 xi Figure 4.4 (a) Phase contrast image of post-annealed bulk GdSiz layer formed on the Si(001) substrate. (b) Selected area diffraction pattern from one grain of the layer with arrows presenting strong diffraction spots from the orthorhombic phase of GdSiz. (c) HRTEM image showing the [100] orthorhombic zone axis from the GdSiz. The atomic lattice symmetry is indicted by the stacking of (001) planes in an aabb manner .......... 55 Figure 4.5 (a) 3D topography STM image of GdSiz nanostructures grown on Si(001) with (b) a cross-sectional line profile, indicated by the line on (a) ................................. 57 Figure 4.6(a) Cross sectional phase contrast TEM image of a GdSi2 nanostructure grown on Si(001). (b) Higher magnification HRTEM image Showing the interface between the GdSiz and Si. (c) FFT of area shown in (b) ..................................................... 58 Figure 4.7 (a) Cross sectional phase contrast TEM image of a GdSiz nanostructure with multiple thicknesses. (b) Higher magnification HRTEM image shows an aabb lattice stacking, consistent with the orth/tet phase of GdSi2 viewed along the [100] zone axis. (c) FFT from area (b) confirms the crystal structure of the orth/tet phase of GdSiz .......... 59 Figure 4.8 (a) An HRTEM image of a GdSiz nanostructure with a wavy top. Two different lattice symmetries are observed in the nanostructure on the two sides of the dashed horizontal line. Above this line an aabb stacking is displayed, and an FFT from area 1 shows a tilted orthorhombic GdSi2[100] zone. Below the line, close to the Si substrate a rectangular symmetry with aaaa stacking is observed. An FFT fiom area 2 shows a hexagonal GdSi2-x[100] zone. (b) An enlarged image from area 3 illustrates the contrast from mismatch dislocations at the interface, as well as a periodic array of strain contrast marked by arrows ........................................................................ 62 Figure 4.9 (a) An STM image of Tm silicide nanostructures on Si(001) with 2.8 ML Tm coverage. (b) A cross sectional profile along the line shown in (a) ........................ 65 Figure 4.10 (a) An STM image of Tm silicide nanoisland with a modulation in height. (b) Cross-sectional profile along the line shown in (a) ............................................ 65 Figure 4.11 A comparison of the projections and corresponding simulated lattice images of the hexagonal Tm38i5 (a and b) and orthorhombic TmSi (c and d) imaged down different axes ........................................................................................ 67 Figure 4.12 (a) HRTEM image Shows a Tm silicide nanostructure approximately 10 nm in width and less than 1 nm in height. (b) Enlarged lattice image shows a rectangular stacking symmetry. (c) FFT shows streaked diffraction peaks ............................... 68 Figure 4.13 (a) Cross sectional HRTEM image of a Tm Silicide nanostructure on Si(001). (b) Enlarged atomic lattice shows rectangular symmetry. (c) Corresponding FFT from area (b). ((1) X-ray energy dispersive spectrum shows chemical nature of the nanostructure ......................................................................................... 70 xii Figure 4.14(a) Cross-sectional HRTEM image of Tm silicide nanostructure from an elongated Tm silicide nanostructure ~90 nm wide and up to 3nm thick between Si substrate and Can layer. (b) Enlarged lattice image shows a stacking similar with the simulated image of orthorhombic TmSi viewed along [100]. (0) Corresponding FFT Shows a TmSi [100] zone. ((1) and (e) shows a comparison of FFT filtered lattice image from the nanostructure and Si, which confirms the stacking in the nanostructure is consistent with the simulated image of orthorhombic TmSi viewed along [100] (inset in (d)) ................................................................................................... 71 Figure 4.15 (a) Cross-sectional HRTEM image of Tm silicide nanostructure on Si(001) shows a stacking symmetry with a half-atom shift between every two rows of atoms consistent with the orthorhombic TmSi [001] axis. (b) and (c) are the corresponding FFTS from areas (1) and (2) ...................................................................... 73 Figure 4.16 Two types of lattice stacking are observed in each of the two nanostructures (a) approximately 25 nm wide and approximately 2nm thick (b) approximately 10 nm wide and approximately 2nm thick. Different stacking are marked by arrows and are compared to the simulated images ............................................................... 74 Figure 4.17 The energy loss near-edge structure (ELNES) of the Gd N edge from metallic Gd, bulk GdSiz, and Gd203 powder in comparison to the Gd203 spectrum from the standard EELS Atlas [27] ........................................................................ 78 Figure 4.18 The energy loss near-edge structure (ELNES) of the Gd M45 edge from metallic Gd, bulk GdSiz, and Gd203 powder in to the Gd203 spectrum fiom the standard EELS Atlas [27]. The line with open squares in (b) demonstrate good curve fitting using a power-law backgound, Hartree-Slater cross section and a set of Lorentzian profiles..79 Figure 4.19 Gd electron configurations for the ground state neutral gaseous atom ........ 80 Chapter 5 Figure 5.1 Phase contrast TEM image of as-sputtered Co/Al multilayers on a Si substrate. Inset is a selected area diffraction pattern from both the multilayers and Si substrates showing strong diffraction from the close-packed growth planes of both the Co and Al, parallel with Si(001). In the upper right, an inclined gain is illustrated schematically to Show that the waviness perpendicular to the electron beam is up to 2 nm at the interfaces. ......................................................................................................... 83 Figure 5.2 HRTEM images showing Co/Al atomic lattices with different contrast. The Co/Al interfaces are poorly defined. A gain boundary in Al is observed in (a). Both FCC and HCP stacking in the Co are Shown in the enlarged images in (b). Insets in the upper right show the corresponding FFTS from the images ................................... 86 Figure 5 .3 A through-focus series of images from the same area of the multilayer indicates that the Al and Co interfaces are inclined along the beam direction .............. 87 xiii Figure 5. 4 Schematic growth model ofFCC A1 {111} on FCC Co W{l 1 l} with a lattice mismatch of 13. 7%. ... .. ... ... ... ... . .. . .......88 Figure 5.5 Phase contrast TEM image of an annealed Co/Al multilayer. Inset is a selected area diffraction pattern fiom both the multilayers and Si substrates showing strong diffraction fi'om the close-packed gowth planes of both the Co and Al ............ 89 Figure 5.6 (a) A dark field STEM image of the as-sputtered Co/Al multilayers showing where the EELS spectrum image was acquired along the white dashed line. The spatial drift square is the cross-correlation scan area used to correct the drift. (b) A comparison of line profiles of image gey level across a Co/Al/Co layers from both as-Sputtered and annealed multilayers. (c) A perspective view of an EELS spectrum image showing the plasmon intensity changing across the Co/Al multilayers ..................................... 91 Figure 5.7 A comparison of the plasmon intensity profile for as-sputtered and annealed Co/Al multilayers. The Al plasmon peak profiles have full width at half maximum of 5-6 nm, similarly for both as-received and annealed samples. The only difference is that the low variation of Co intensity length is 15 nm for the as-received multilayers and 17 nm for annealed one92 Figure 5.8 A perspective view of the Co white line Spectrum image scanned across Co/Al/Co layers over a length of 20 nm of the as-sputtered multilayers ..................... 94 Figure 5.9 Co white line intensity profiles for both the as-sputtered and annealed multilayers. A theoretical profile simulated by using a 0.5nm FWHM Gaussian electron beam profile across perfect Co/Al interfaces is used to compare with the experimental data .................................................................................................... 94 Figure 5.10 Phase contrast TEM image of as-sputtered and annealed Py/Al multilayers. Insets are selected area diffraction patterns from the multilayers. Both morphologies and diffraction patterns show similar structures in two conditions. 96 Figure 5.11 HRTEM image of an annealed Py/Al multilayer along a Py <110> zone with corresponding FFT in the inset showing the gowth relationship .............................. 97 Figure 5.12 A comparison of plasmon intensity profile from the as-sputtered and annealed Py/Al multilayers. The Al plasmon peak profiles have full width at half maximum of 6.5-7 nm, Similarly for both as-sputtered and annealed samples. . . . . .....98 Figure 5.13 Ni (a) and Fe (b) white line intensity profiles for as-sputtered and annealed multilayers are compared separately with the theoretical profile simulated by using a 0.5nm FWHM Gaussian electron beam profile across perfect Py/Al interfaces. (c) A comparison of line profiles of the image gey level from Py/Al/Py layers in the Z-contrast images of the two conditions ......................................................... 99 xiv Figure 5.14 Phase contrast TEM image of as-sputtered Cu(8nm)/Al(10nm) multilayers. Observed layers are in different contrast, exhibiting ~12nm thick Cu-rich layer and ~6nm thick Al-rich layer, which are different fi'om the nominal thickness ....................... 101 Figure 5.15 HRTEM image shows three types of stacking fiinges with different contrast existing in the Cu/Al multilayers. Five different areas in the image are analyzed using FFTS, resulting in three crystal zone patterns shown in the insets. Areas 1 and 5 show the same 0 AlzCu <001> zone patterns, areas 2 and 4 Show the same [3 AlCu3 <111> zone patterns, and area 3 shows a 0 AlzCu <001> zone pattern .................................... 103 Figure 5.16 Schematic representation of the bee AlCu3 and tetragonal A120u crystal structures and the orientation relationships of their close-packed planes. (a) bee AlCu3 crystal structure (b) tetragonal AlzCu crystal structure (c) a structural rectangle unit in AICU3(110) ((1) two orientation relationships with A12Cu(l 10): AlCu3[111] // A12Chr[001] (the fine dotted lines) and AlCu3[111] // A12Cu[111] (the coarse dotted lines). .....106 Figure 5.17 Phase contrast TEM image of annealed Cu(8nm)/Al(10nm) multilayers. The Cu-rich layers are ~13 nm thick and the Al-rich layers are ~7 nm thick ................... 107 Figure 5.18 HRTEM image with corresponding FFTS from the armealed Cu(8nm)/Al(10nm) multilayers. The Al-rich phase and the Cu-rich phase are identified to be A12Cu and AlCu3 respectively ............................................................. 107 Figure 5.19 Phase contrast TEM image of the as-sputtered Cu(5nm)/Al(3nm) multilayers. The Cu-rich layers are 3-4nm thick and the Al-rich layers are 4-5 nm thick .............. 109 Figure 5.20 HRTEM image shows two types of lattice fringes existing in the Or(5nm)/Al(3nm) multilayers. Four different areas in the image are analyzed using FFTS, resulting two crystal zone patterns shown in the insets. Areas 1 and 3 show the same 0 A12Cu <111> zone patterns, while areas 2 and 4 show the same Cu <110> zone patterns .............................................................................................. 109 Figure 5.21 Schematic representation of FCC Cu (111) and tetragonal AlzCu (110) gowth orientation and lattice mismatches ..................................................... 110 Figure 5.22 TEM bright field image of an as-Sputtered spin-valve with nominal thickness in nanometers Nb(1 50)Cu(5)FeMn(8)Py(24)Cu(l 0)Al(10)Cu(1 0)Py(24)Cu(5)Nb(50) .................. 1 1 2 Figure 5.23 Phase contrast TEM image of Py/Cu/Al/Cu/Py layers in the as-sputtered spin-valve with x=10nm ........................................................................... 1 13 Figure 5.24 (a) HRTEM image of Py/FeMn/Cu/Nb layers in the as-sputtered spin-valve with x=10nm and (b) the corresponding FFT from the image .............................. 114 XV Figure 5.25 (a) HRTEM image of Cu/Al/Cu layers in the as-Sputtered spin-valve with x=10nm and (b) corresponding F FT from the image ....................................... 116 Figure 5.26 (a) STEM bright field image of the multilayers in the as-sputtered spin-valve. (b) XEDS intensity profiles of Cu, Ni, Fe, and Al were collected fiom the line across the Py/Cu/Al/Cu/Py/FeMn/Cu layers in the image (a) .......................................... 117 Figure 5.27 TEM bright field image of an annealed spin-valve with nominal thickness in nanometers Nb(150)Cu(5)FeMn(8)Py(24)Cu(10)Al(30)Cu(10)Py(24)Cu(5)Nb(50)...l 19 Figure 5.28 Phase contrast TEM image of Py/Cu/Al/Cu/Py layers in the annealed spin- valve with x=30nm .............................................................................. 120 Figure 5.29 HRTEM image of the Al/Cu/Py layers in the annealed spin-valve with x=30nm. Five different areas in the image are analyzed using FFTS. Areas 1 shows a 0 A120u <100> zone pattern, area 2 shows a B’ AlCu3 <111> zone pattern, area 3 shows a B’ AlCu3 <001> zone pattern, area 4 shows a BCC Py <111> zone pattern, and area 5 Shows a FCC Py <111> zone pattern ......................................................... 122 Figure 5.30 (a) STEM bright field image of the multilayers in an annealed spin-valve. (b) XEDS intensity profiles of Cu, Ni, Fe, Al and O are collected from the line across the Py/Cu/Al/Cu/Py layers in the image (a) ....................................................... 124 Images in this dissertation are presented in color. xvi Chapter 1 Overview Great achievements have been obtained in nanotechnology in the last decade by taking advantage of the unique properties of materials structured on the nanometer scale (normally between 1-100 nanometers). Nevertheless, due to lack of fundamental understanding of the thermodynamic and kinetic characteristics of nanoscale materials, there is still a large barrier between rational desigr and controlled synthesis with desired properties at the nanometer scale [1]. It is well known that bulk material properties are not critically dependent on Size, but the properties of nanoscale materials are often very size-dependent. Consequently, the structure, property, and stability of nanoscale materials must be understood to enable their desigr and application. The objective of this work is to study the gowth and crystallogaphy of a number of nanoscale heterostructures by characterizing their crystal structures, interfacial epitaxy, and interfacial chemical and electronic nature. This will provide basic understanding of the thermodynamic phase stability and surface energies that affect nucleation, gowth, and stability. The selected systems in this thesis are two classes of nanoscale electronic material systems. The first type of materials systems are epitaxial rare earth Silicides, which can self-assemble to form nanowires on silicon(001) substrates under specific gowth conditions. Different morphologies, nanowires and nano-sized islands, were observed during the gowth of the Silicides. The morphologies are correlated with the different crystal structures of the RE Silicides, which have different lattice mismatches with the Si substrates. The second type of materials systems are sputtered magnetic multilayers, including giant magnetoresistance (GMR) multilayers and spin-valves, which are the basic components for many read heads in hard-disk drives. These nanometer thick metallic multilayers show interesting magnetoresistance (MR) behaviour, such as large current-perpendicular-to-plane (CPP) specific resistances, small interface scattering asymmetries, and unstable MR in terms of time and temperature. These interfacial properties and the changes in MR may be affected by individual layer structures, interfacial roughness, and/or the thermodynamic stability of the structures. High resolution transmission electron microscopy (HRTEM) has been carried out to investigate the crystal and interfacial structures of these nanoscale electronic materials. The phases formed and epitaxial relationships between the phases have been characterized to obtain an understanding of the gowth mechanisms and stability of the structures at the nanometer scale. Simultaneously, analytical electron microscopy, including electron energy-loss spectrometry (EELS) and x-ray energy dispersion spectrometry (XEDS), have been used to investigate the chemical and electronic nature of these structures, which can provide useful information for understanding their physical and chemical properties. Chapter 2 Background This chapter deals with the current critical backgound literature for this thesis. It will start with an introduction of the fundamental theory of high-resolution transmission electron microscopy (HRTEM) and electron energy-loss spectrometry (EELS). This is followed by a literature review of the fundamentals of the epitaxial rare earth silicide nanowires and giant magretoresistance multilayers. 2.1 Fundamentals of HRTEM and EELS Transmission electron microscopy (TEM) has been demonstrated to be an efficient technique for investigating the crystal structures and chemical nature of materials. TEM is becoming more important as the dimensions of many material structures approach the nanometer scale, which often lie below the spatial resolution of other experimental methods. Hi gh-resolution transmission electron microscopy (HRTEM) is a particularly powerful technique for the investigation of local crystal structures on the atomic scale and is very useful for understanding the details of nanoscale materials. Complemented with analytical capabilities, including x-ray energy dispersive spectrometry (XEDS) and electron energy-loss spectrometry (EELS), HRTEM’S spatial resolution in the nanometer and sub-nanometer range, makes it an indispensable tool for nanoscale analytical application. In this thesis, HRTEM and EELS are the primary techniques being applied to the study of two classes of nanometer scale electronic materials: self-assembled rare earth silicide nanostructure epitaxial gown on Si(001) substrates and giant magretoresistance (GMR) multilayers and spin-valves. 2.1.1 HRTEM TEM image contrast arises from the scattering of the incident electron beam by the specimen. Both the amplitude and phase of the electron wave can be changed, which gives rise to image contrast. Conventional TEM, including bright field (BF) and dark field (DF) imaging, takes advantage of diffraction contrast (one type of amplitude contrast). However, high-resolution TEM is essentially phase contrast imaging. While BF or DF imaging selects a single beam to form an image, a phase contrast image contains more than one beam. The contrast observed can be described in a quantitative way that breaks the process into a number of steps. These include a specimen function and contributions fi'om the microscope. 2.1.1.1 Specimen Function The specimen interacts with the incident electron beam because of the electronic static potential of the specimen. Therefore, there is a phase shift of an electron when passing through a specimen, which is 0 = other (2.1) where o is the interaction constant, (Do is the mean electrostatic potential, and r is the thickness [2]. The interaction of the incident electrons with the specimen is described with the specimen transmission function f(r), 11w) = eXPl-i6¢o(X.y)-u(x.y)]. (2.2) where (Do(x,y) is the projected potential of the specimen taking account of variations in the z-direction and u is an absorption term [2]. If the specimen is thin, the absorption part can be igrored. The resulting wave exiting the specimen is described as WAX,” = f(X,Y) \P0(X’Y)' (2'3) which means the amplitude of a transmitted wave function (‘11,) will be linearly related to the projected potential of the specimen for a very thin specimen. This is also called weak-phase-obj ect approximation. 2.1.1.2 Microscope Effect After the electrons leave the specimen, the ability of a lens to pass the electrons is described with the transfer fimction T(u): T(u) = 2A(u) sin x(u) (2.4) where u is a reciprocal space lattice vector, A(u) is the aperture function, and x(u) is the phase—distortion function, which is expressed as a function of the spherical aberration (Cs), defocus (AI), and electron wavelength (A): x(u) = niAfu2+1/2nc,ru“ (2.5) Plots of x(u) vs. II, and T(u) vs. u are shown in figure 2.1a and 2.1b [2]. The best resolution that allows direct interpretation of bright/dark contrast as atom position corresponds to the first zero point of T(u), or sin x(u), which is called the Scherzer resolution, as indicated in figure 2.1. The Scherzer resolution is achieved when the transfer function is optimized by balancing the effect of spherical aberration at a particular negative value of Af, which is known as Scherzer defocus: %...... = -1-2 0.1 (2.6) At this defocus condition, the electrons at different reciprocal lattice positions are transmitted in a nearly constant manner (similar phase and amplitude) up to the Scherzer resolution, which is 0.66Cl/4/I3/4 [2]. The TEM used for HRTEM in this study is a JEOL JEM2200F S working at 200KV, which has a Cs of 0.5 mm, and hence, a theoretical point-to-point Scherzer resolution of 0.19 nm. Note however, that information at larger reciprocal space frequencies (finer real space resolutions) can be passed. Imaging at such higher frequencies allows imaging structures below the Scherzer limit. Nevertheless, image interpretation becomes more difficult with such “pass ban ” imaging, because not all information is passed in the same manner. The ultimate resolution is limited by the spatial coherence spread of the electron source and chromatic aberration. Hence, T(u) is damped by imposing an envelope function on it, as Shown in figure 2.1. It is evident that T(u) is not able to transfer any information with u larger than a specific value. This resolution limit is known as the information limit, as indicated in the figure. (3) 511196; 098 oar/I III >u ..-; _____ =\llll \V II (b) 2 Scherzer resolution T(u) \\information limit envelope function Figure 2.1 (a) sin x(u) vs. II (without damping) at the Scherzer defocus, and (b) T(u) vs. II modified by imposing an envelope function (dashed line); Af = -100nm, C5 = 2.2 mm [2]. 2.1.1.3 Fast Fourier Transform Fast Fourier transform (FFT) analysis is an effective way to interpret the atomic structure in the HRTEM images. F FT diffractogams are obtained by capturing an area of interest in the images. The atomic lattice spacings and symmetries (including interpretable allowed and/or forbidden reflections) can been determined from F FT, which in turn provides useful information for analyzing the crystal structures. Such FFTS can be used in a manner analogous to selected area diffraction (SAD) patterns, without the spatial limitation of SAD patterns [2]. Figure 2.2 shows an HRTEM image of single crystal Si viewed along Si[110] and its corresponding FFT diffractograrn. Calibrated for the appropriate image magnification, these Si FFT spot spacings represent the reciprocal lattice spacings of the Si crystal planes of the imaged zone (i.e. Si {1 1 1} and Si{002} for the <110> zone axis shown here). Note that additional effects, such as spot spreading can result from edge effects. Z.‘ 'iij.‘.l.';l 'suTIn ‘an” ’ "' _ ~ R )1 Si(lTl) ' ‘~'Sil00l‘l‘ -Si(l)02) Figure 2.2 (a) An HRTEM image of Si along [110] zone (b) corresponding FFT pattern. The reciprocal spacings of the spots are related to the corresponding Si atomic planes. 2.1.2 EELS 2.1.2.1 Fundamentals Electron energy-loss spectrometry (EELS) is based on the general idea that as incident electrons interact with a Specimen, some electrons will lose energy (inelastic scattering). Specifically, the primary electron loses some energy by transferring it to the material, thereby exciting electrons in the specimen to higher energy states. Measurement of the energy distribution of these inelastically scattered electrons leads to the so-called electron energy-loss spectrum (EEL spectrum). These Spectra have been used to analyze the electronic structure of specimen atoms, which in turn reveals details of the nature of these atoms, their bonding and nearest-neighbor distributions, and their dielectric response [3]. Figure 2.3 shows a typical EELS spectrum from an NiO Specimen. The most intense feature seen in the spectrum is the zero-loss peak. This peak is generated from elastically scattered electrons. It has a certain width (1-2 eV) caused by the non- monochromaticity of the primary electrons and some instrumental broadening. The plasmon peak, which occurs because of oscillation of weakly bound electrons, is the second most dominant feature of the EEL spectrum after the zero-loss peak. The other features are composition related edges due to the electron excitation in the specimen to specific states, which can be used to study the chemical as well as the electronic nature of the materials. «Zero Loss O K edge A/ gain change Plasmon Loss Intensities (arb. units) A/ NI M edge I ' I I n 1 n 0 200 400 600 800 1000 Energy Loss (eV) Figure 2.3 A typical BEL Spectrum from an NiO specimen including zero-loss peak, plasmon peak, and compositional edges for NiO. EELS edges are indicated by the letters K, L, M, N followed by one or more numbers. The letter indicates the atomic shell from which the incident electrons excited bound electrons, while the number indicates the orbital. Table 2.1 gives an overview of the most important edges. For example, the L1 indicates an excitation of a 25 state, L2,; indicates an excitation of 3 2p state, which is spin split into 2pm (L2) and 2pm (L3) states. The spin split states are labeled by their total angular momenttmr J and are (21+1)-fold degenerate, e.g. a 2pm state can have j=—3/2, -1/2, 1/2, 3/2 and is therefore 4—fold degenerate [2]. The degeneracy of a state determines the number of electrons that can be in that state and is important because the more electrons there are in a state, the more likely an excitation from that state will be. As a result, one would expect that the intensity ratio between an L3 and an L2 edge to be 2:1 [4]. Table 2.1 Important edges in EELS associated with their electronic states excitation and degeneracy. Edge State Degeneracy K 1 s1 ,2 2 L1 251,2 2 L2 2131/2 2 L3 2133/2 4 M1 331,2 2 M2 3pm 2 M3 3133/2 4 M4 3d3,2 4 M5 3d5,2 6 2.1.2.2 Energy-Loss Near-Edge Structures in EELS The energy resolution of EELS (ranging fi'om approximately 0.3 to 3 eV, depending on instrument) is much better than X-ray spectrometry (approximately 135 eV for XEDS). Because of this, EELS can be used to determine a wealth of information about the Specimen in addition to its basic elemental chemistry. Much of this information is contained in fine-detail intensity variations in the ionization edges, termed energy-loss near-edge structure (ELNES). From these fine structures, information such as the bonding state of the ionized atom, the coordination of the atom, and the density of states, can be obtained. For example, due to different atomic bonding between gaphite (Sp2 bonds in the basal plane with van der Waals bonding between the planes) and diamond (four directional hybridized Sp3 covalent bonds) [5], one can easily distinguish gaphite from diamond from the carbon K edges in the EELS Spectra. 10 EELS spectra Show intense peaks for L2; edges of transition metals and M4; edges of RE metals. These spectral features are called “white lines” because of their appearance on the photogaphic plates in early experiments with x-ray absorption spectroscopy (XAS) [2]. The positions and relative intensities of these edges can be influenced by local chemistry and the bonding between atoms due to the solid state effect (i.e. the inelastic scattering of electrons is affected by the atomic crystal periodicity rather than a single atom [6]). Leapman et a1. [4] have studied these fine structures of 3d transition metals and their oxides using EELS. They found that there is a difference in the L3 threshold energy (chemical shift) between the metal and oxide. Furthermore, the statistical ratio of intensities of L3 and L2 has been found to deviate fiom the statistical value (2.0) based on the occupancy of the 3d states, i.e., with values ranging between 0.8 for Ti and 5.0 for Fe. In the case of metallic copper, the firll 3d states lead to the absence of threshold peaks, while in compounds such as CuO electrons are drawn away from the copper atom, leading to d-level being empty and sharp L3 and IQ threshold peaks. Kurate et a1. studied EELS of a series of Mn oxides and demonstrated that the intensity ratio of the white lines is related to the charge state or valency of the elements [7]. ll 2.2 Self-Assembled Silicide Nanowires The epitaxial gowth of rare-earth (RE) metal Silicides on Si(001) has attracted interest Since nanowires have been found to self-assemble on Si substrates due to an anisotropic lattice mismatch [8-14]. These nanowires display metallic conduction [10, 14] and consequently are promising candidates for firture nanoscale interconnects and devices. In addition to this technological interest, detailed study of the crystal structures, defects, and stoichiometry at the interface of the nanostructures and Si surface should give a basic understanding of epitaxial thin films in lattice mismatched systems. The motivation of this study is to investigate the gowth mechanisms and electronic structure of self-assembled RE silicide nanowires using HRTEM and EELS. In the following sections, backgound and basic information on RE Silicides are introduced, along with more detailed material on the electronic nature of these nanowires. 2.2.1 Introduction to Self-Assembled Silicide Nanowires In 1998, Preinesberger et al. [8] first reported the gowth of RE disilicide nanowires formed by evaporating a small amount Dy on to a Si(001) surface followed by subsequent annealing at high temperature. Later Chen et a1. [9] explained that the self- assembly of Er disilicide nanowires on Si(001) is due to the anisotropic mismatch between the hexagonal Ale type structure of the Silicides and the Si(001) surface. The nanowire is free to gow along ErSiz [1120 ], which has a small mismatch (-1.3%) with the hexagonal a axis, and is constrained from gowing along ErSiz [0001] where there is a large mismatch (+65%) with the hexagonal c axis. Besides Dy and Er, other RE metal Silicides including Sm [l3], Gd [11, 14], Ho[10,12], Yb [15], Se [1 1], Y [16] and Tm l2 [l7] Silicides are reported to form nanowires on Si(001). A summary of the lattice constants of RE Silicides and their gowth behavior on Si(001) is shown in Table 2.2. Although Y and Se are technically not rare-earth metals, their Silicides have similar properties as the lanthanide metal Silicides, so typically Y and Sc Silicides are also included as RE Silicides. Table 2.2 A summary of the lattice constants of RE Silicides and their gowth behavior on Si(001). Lattice parameter (mismatch Silicides %) Nanowires a (M!) (%) 6 (nm) (%) ScSi1'7[1 1] 0.366 (-4.7) 0.387 (+0.8) Yes YSi1_7[l6] 0.384 (0) 0.414 (+7.9) Yes NdSil.7 [18] 0.412 (+7.3) 0.444 (+16) No Sm3Si5 [13] 0.390 (+1.6) 0.421 (+9.6) Yes GdSi 1'7 [11] 0.389 (+1.0) 0.417 (+8.7) Yes DySi 1.7 [8] 0.383 (-0.2) 0.412 (+7.3) Yes HoSi 1.7 [10] 0.382 (-0.6) 0.411 (+7.0) Yes ErSi 1.7 [9] 0.380 (-l.3) 0.409 (+6.5) Yes Yb3Si5 [15] 0.378 (-1.5) 0.410 (+6.7) Yes Tm3Si5 [17] 0.377 (-1.9) 0.407 (+6.0) Yes 2.2.2 Crystallography of Epitaxial Silicide N anowires 2.2.2.1 General Understanding of Bulk RE Silicide Phases Before going into the details of RE Silicide nanostructures, it is very important to gain a general understanding of the known bulk phases of the RE Silicides. In many cases, the phase equilibrium of bulk RE Silicides are poorly defined, and there is 13 considerable uncertainty regarding the solubility and temperatures where the RE Silicides exist. The Gd-Si phase diagarn (figure 2.4) and crystal structure data of the system (Table 2.2) [19] show that 7 phases can form depending on the temperatures and stoichiometries. Since in the gowth of nanostructures, only a small amount of the RE metals are evaporated onto Si substrates, and the substrate temperature is typically around 600°C (see Chapter 3.2), the Si-rich Silicides are expected to form. On the Si—rich side of the Gd-Si phase diagam, GdSi2 forms tetragonal or orthorhombic phases over a narrow stoichiometry range depending on temperature, while at a slightly different stoichiometry, GdSi2.x forms a poorly defined [3 phase at high temperature and a hexagonal phase below 1000°C. Consequently, any of these phases with similar stoichiometry might form the RE silicide nanostructures on Si substrates. 2100°C 2100' L 9 ~ - it ~ 2' ‘ ~ 9 'I' ' "‘7' \\ 1800- o 1. .t - - - ‘ _ 1650120_C ”3;: a ‘. _ 1313°C ‘. ."u- ..«S ‘ _ 1500 ’ I .‘ : - ‘3' . g \\ 1414OC‘ o KTIBGd) 1’ ' ‘5 ~1200°C . °~ 1200'\ 1’ 1070°C .. fl. """" ~ 735 """ '- ' ? g \ o 15.3 ? a 900- 1235 C 1060120°C T (._BGdSi2 N E 600‘ m .3 _ E o - «.0... mm 3.. g .. ”aegis _____ 0 60 ’64668 (Si)-> 0 2'0 4'0 60 8'0 1(0 Gd Si Atomic Percent Silicon Figure 2.4 Gd-Si phase diagam, after [19]. 14 Table 2.3 Gd-Si crystal structure data [19]. Pooooo Gd5Si3 37.5 P63/mcm MnSSi3 hexagonal Gd5Si4 44.5 P41212 Ge4Sm, tetragonal GdSi 50 ana FeB orthorhombic terSi2_x 60 to 62.5 P6/mmm AlB2 hexagonal [3GdSi2_x 60 to 62.5 0tGdSi2 64 to 66.8 Imma GdSi2 orthorhombic [3GdSi2 64 to 66.8 I4lamd ThSi2 tetragonal Many RE-Si systems (such as Dy-Si, Y-Si) have similar phase diagams as that of Gd, with hexagonal, orthorhombic, and tetragonal phases being reported around the stoichiometry of RESi2 [19]. Unlike Gd and most RE Silicides, which have three possible disilicide phases with very close stoichiometry, the Tm-Si phase diagam (figure 2.5) predicts three phases with very different stoichiometries [19]. Among the three phases Tm5Si3, TmSi, and Tm38i5, the latter two are phases that may form in a Si rich enviromnent. 15 0 L 1600 ‘é/l545 C 9 9 9 l4l4°C Do 1200- ? """" .7" """ - 2 E E" 800- .. D. E h M In 400- :7) 3;) ii ' é” E E H I" I- 0 . . . . 0 20 40 60 80 100 Tm Si Atomic Percent Silicon Figure 2.5 Tm-Si phase diagam, after [19]. 2.2.2.2 Self-Assembled RE Silicide N anostructures The formation of nanowires is attributed to the anisotropic mismatch between the hexagonal Ale type RE Silicides and the Si substrate, as illustrated in figure 2.6. The silicide nanowires are proposed to gow with the silicide [0001] and [11 20] directions parallel with the (001) surface, lying along the two perpendicular Si <110> directions of the substrate. Using Dy silicide as a prototype for the RE Silicides that form nanowires on Si(001), hexagonal DySiz, with lattice constants a=0.383nm and c=0.412nm, corresponds to mismatches with the Si <110> of -0.2% and 7.3% respectively. As a result, nanowires gow to arbitrary length in the direction of small mismatch, but are limited in their ability to gow coherently with the substrate in the lateral direction [9]. 16 a C NW growth Figure 2.6 Schematic representation of a hexagonal Silicides forming a nanowire on Si(001) due to the small mismatch along the 3 axis and large mismatch along the c axis, after [20]. Besides nanowires, nanostructures with other morphologies, such as nano-size islands, can be observed during silicide gowth (as seen in figure 2.7). It is intuitive to imagine that either tetragonal or orthorhombic Silicides contribute to this type of morphology. However, these may form with different stoichiometries. Hence, identifying the silicide structures and relating them to the nanostructure morphology is important in understanding the gowth mechanisms of the epitaxial Silicides. IRTEM, combined with fast Fourier transform (FFT) analysis, is an effective way to identify the structures of these nano-size features. Using both cross-sectional and plan view HRTEM, Ye et al. demonstrated that both hexagonal and orthorhombic (or tetragonal) structures exist when Dy Silicides gow epitaxially on Si(001) [21]. 500 nm Si [100] nanoislands %.‘I' \. 4. Si [010] \. . v/ I] a n 0“" l I‘CS r/ a» 0 500 nm Figure 2.7 An atomic force microscope (AFM) image of GdSi; nanowires and nanoislands gown on Si(001) substrate along two perpendicular Si <110> directions, after [17]. It is generally accepted that epitaxial gowth occurs when the lattice constant of the epilayer closely matches that of the substrate. The lateral gowth will be limited only if there is a sigiificant lattice-mismatch (>2%) between the epilayer and the substrate [11]. For those Silicides that can form nanowires on Si(001), the magnitudes of lattice-mismatch in the smaller mismatch direction are all within i2%. Figure 2.8 shows the magnitudes of the anisotropic mismatches in two directions for hexagonal Ale type of RE silicide nanowires with respect to Si<110>. Different magnitudes of lattice mismatches for these Silicides may result in different ways to relax the strain (such as dislocation generation and/or epilayer tilt) between the epilayer and the substrate. Investigating the interfacial crystallogaphy by HRTEM will reveal important information for understanding the gowth mechanisms of the RE silicide nanowires. 10 _ : . . _ Sm3Si5 Q 8 _ Yb3Si5 . o +vsrz GdSiz 9.. 6 _. . O . DySi' ‘ d: . ErSiz "”512 I 3 PTm3SIs I I a 4 P | a I .2 2 .. I 2 . : ‘9 0 . ...................... ' ...................... '3 - :' a - .. A 2 _ E 0 PE” "‘- i - ..1 -6 - : ScSiz i. I -3 _ I I _10 L l . l . l I l 1 l . l -2.0 -l.5 -l.0 -0.5 0.0 0.5 1.0 1.5 2.0 Small Lattice Mismatch (%) Figure 2.8 Plot of hexagonal AlB2 type of RE silicide anisotropic mismatches with Si(001). With respect to Si(001) two perpendicular <110> directions, the lattice mismatches of Silicides have a large absolute value between 4% and 10% and a small lattice mismatch within :l:2%. (after [1 1]) Among the RE Silicides, it is especially interesting to note that hexagonal Tm silicide has a large lattice mismatch (-l .9%) in the small mismatch direction, very close to the generally accepted limit (2%) for free epitaxial gowth [l 1]. Recent work has shown that Tm Silicides form epitaxial nanowires on Si(001) [17]. Nevertheless, there are also islands present. An AFM image Shown in figure 2.9 shows the typical morphologies of Tm silicide nanostructures: nanowires with high aspect ratio and islands, typically much lower aspect ratios, are often observed at nanowire intersections. In addition, there is a very low density of large bright features may be clusters of silicon l9 carbide. These features are both much wider and much higher than the silicide related features. The different silicide morphologies might be correlated with different crystal structures such as the Ale hexagonal Tm3Si5 (TmSiL7) and BCr orthorhombic TmSi. In this thesis work, cross-sectional HRTEM was carried out to identify the crystal structures of these Tm silicide nanostructures. It is also interesting to investigate the relaxation of the strain between the epilayer and the substrate, since compared to other RE Silicides, a larger magritude of lattice mismatch must exist in the Tm silicide nanowire gowth direction. Furthermore, due to the potential large differences in stoichiometry of the Silicides associated with these two morphologies, it may be that transformation between morphologies might be more difficult than in the other RE silicide systems, which may result in the potential to form more stable nanowires. unknown clusters II II L l/ Vi- ' , nanowires 500 nm — Figure 2.9 An AFM image of Tm silicide nanostructures on Si(001) with 0.98 monolayer (ML) Tm coverage. 20 2.2.2.3 Crystal Structure of Bulk Phases in the Gd-Si and Tm—Si Systems Three potential Gd Silicide phases, including the hexagonal, tetragonal and orthorhombic phases, may form the silicide nanostructures on Si(001) substrates. Figure 2.10 shows crystal models of the prototype hexagonal Ale structure of GdSiz.x (a) and the prototype tetragonal ThSiz structure of GdSi; (b). The prototype orthorhombic GdSi2 structure is a slight distortion of the tetragonal ThSiz structure, with a slight variation of lattice parameters a and b. However, since it is difficult for TEM to identify the slight lattice differences between the orthorhombic and tetragonal phases, orth/tet will be used to describe these silicide phases in this thesis. The lattice parameters of these phases and their lattice mismatches to Si<110> are listed in table 2.4. The hexagonal phase has an anisotropic mismatch with Si, which may result in the NW formation. Both of orthorhombic and tetragonal phases have large lattice mismatches in two perpendicular directions, which suggests they will not form nanowires but instead form nanoislands. c—axis [001] . RE 0 Si c-axrs a-gxis fig b-axis [00011 [211°] [100] [010] (a) hexagonal structure (b) orthorhombic or tetragonal structure Figure 2.10 Crystal models of RE Silicides, (a) hexagonal Ale structure (b) orthorhombic GdSiz or tetragonal ThSiz structure. (after [20] ) 21 Table 2.4 Gd disilicide phases and lattice mismatches with Si<110>, after [19]. Struct a b c Mismatch(%) ure (Hm) (nm) (nm) a b c GdSi,_x hexagonal 0.388 0.417 1 8.6 orthorhombic 0.409 0.401 1.344 6.51 4.43 GdSi 2 tetragonal 0.410 0.410 1.361 6.77 6.77 To study the crystal structures using cross—sectional HRTEM, it is useful to examine the different phases projected down various axes. The crystal structures of the hexagonal and orthorhombic phases, together with the projections along different axes, are shown in figure 2.11. The projection along the hexagonal a axis shows rectangular symmetry with aaaa stacking of the strong scattering RE cations, quite different from the symmetry seen along the orthorhombic a or b axes, which have aabb stacking of the RE atoms. The projection along the hexagonal c axis shows 6-fold symmetry. The different symmetries in the projections along different axes can help to distinguish the crystal structures. (a) “0 Hexagonal GdSi“ ‘ . o 0 g ‘ . O . O Orthorhombic GdSi2 . . O o O O O . . G G 0 e e e i : z e . 0 ¢ 0 G O 0 I o 0 Along hexagonal ‘ . 0 . o \ an axis (or [100]) . O K O O O . O . O c O 0 Along orthorhombic .°.'.°g aorbaxis(or[100]) O O O 0 0 00000 A! O 0 O ‘ G_d Along hexagonal ::::: . Sr c axis (or [001]) . Along orthorhombic c axis (or [001]) Figure 2.11 Projections along various axes of the hexagonal and orthorhombic GdSi; phases. 22 Three Tm silicide phases, including hexagonal TmSSi3, orthorhombic TmSi, and hexagonal Tm3Si5, and their lattice mismatches to Si<110> are listed in table 2.5 [19]. Both the AlB2 hexagonal Tm3Si5 (TmSim) and BCr orthorhombic TmSi have reasonable anisotropic mismatches with Si<110> and are candidates to form the nanowires on Si(001). Figure 2.12 shows crystal models of the prototype hexagonal Ale structure (a) of Tm38i5, the prototype tetragonal CrB structure of TmSi (b) and the projections of the two phases down various axes. The projection along the hexagonal Tm3Si5[001] shows 6-fold symmetry, which is very similar with the projection along the orthorhombic TmSi[lOO]. The projection along the hexagonal Tm3Si5[100] Shows rectangular symmetry with aaaa stacking of the cations, quite different with the symmetry projected along orthorhombic TmSi[001], which has aabb stacking. The difference of the latter two projections is useful for distinguishing between the two phases. Table 2.5 Tm silicide crystallogaphic data and lattice mismatches with Si<110> after [19] Composition Space Crystal Lattice parameters Phase at.% Si goup Prototype structure (mismatch %) = + TmSSi3 37.5 P63/mcm MnSSi3 hexagonal ac=00.862158((+1610l98)) a= 0.418 (+8.9) TmSi 50 Cmcm CrB orthorhombic b=1-035(+170) c= 0.378 (-l.6) Tm3Si5 62.5 P6/mmm Ale hexagonal :: 84377 ((35%)) 23 (a) (b) 828:: f. f. f 0:0:0 Hera onaleS' g "5 0.. 0.0.0 000.0 0 O O R [Along TmJSisc ‘0‘“: axis (or [001]) Along TmSi c axis (or [001]) ”0 0° 0 °.° K 0 ° 0'. o O 0 O O O c ' 3 e e e O 0 Z ° 9 o o e onng,Srsa 0 0 axis (or [100]) ....0 O‘.°. AlonngSia axis(or[100]) Figure 2.12 Projections along various axes of the hexagonal Tm3Si5 and orthorhombic TmSi phases. 2.2.3 Electronic Structure of Silicide Nanowires Epitaxial thin fihns of these RE Silicides gown on Si are reported to have metallic conductivity [10, 14] and form contacts to n-type silicon with low Schottky barrier height [22]. Consequently, these Silicides have been proposed for applications in microelectronic devices and interconnections [23]. For the one-dimensional RE silicide nanowires being considered for electronic applications, it is critical to understand if their electrical and electronic behavior varies between the bulk (or thin film form) and nanophase. Research on the electronic properties of Dy silicide nanowires [24] and Gd 24 silicide nanowires [23] has been performed using angular resolved photoelectron spectroscopy measuring a uniformly oriented array of RE silicide nanowires on vicinal Si(001). The results Show that both Dy and Gd silicide nanowires have one-dimensional metallic character. However, the electronic structure of individual RE silicide nanostructures has not been well characterized. 2.2.3.1 EELS Study of RE Electronic Structure The M4; edge of rare earth metals in the electron-energy—loss spectrum are dominated by the transition of 3d electrons to unoccupied 4f states [25]. To study the near edge fine structures, including edge positions and intensity ratios, it is important to understand the electronic structures of the RE metals. A systematic experimental study of M45 RE oxide edges done by Manoubi et a1. [26] showed that the separation of M5 and M4 peaks is prominent and varies between l6eV for lanthanum and 45eV for ytterbium (Table 2.6), reflecting the Spin-orbit splitting of the initial states in the transition. The white line intensity ratios of RE oxides vary depending on their unoccupied 4f states. It is interesting to investigate if the ELNES of the RE metals varies with their oxides. To date, there have been no published results in the literature or standard EELS Atlas [27]. 25 Table 2.6 White lines energy differences and intensity ratios for the investigated rare earth oxides [26]. Rare earth A (Ms' 4) (eV) I (M5)3I(NI4) La 16.7i0.3 0.76 Ce 18.5i0.2 0.85 Pr 19.6i0.1 1.20 Nd 20.9i0.1 1.45 Sm 25.6i0.1 1.84 Gd 31.2i0.7 1.73 Tb 33.1 i0.4 2.30 Dy 36.7i0.5 4.05 Ho 38.4i0.9 4.92 Er 41.6;I:O.9 5.80 Yb .. oo Lu - It is known that thin film of RE Silicides on Si(lOO) or Si(l 11) display classic metallic behavior [28,29]. As a result, it is interesting to compare the ELNES of the RE silicide thin film with that of the RE metals and their oxides. Furthermore, by comparing the ELNES of M45 edges of the RE silicide nanostructures to those from the bulk materials, the electronic nature, local chemistry, and coordination information of individual nanostructure can be well understood. 26 2.3 Giant Magnetoresistance (GMR) Multilayers This section will introduce some general backgound on giant magietoresistance (GMR) multilayers. It starts with basic concepts related to the GMR effect, and then includes a literature review on the correlation between the GMR effects and multilayer structures. This is followed by the motivation of this study. 2.3.1 Introduction of the GMR Effect GMR is a quantum mechanical effect observed in multilayered structures composed of alternating ferromagietic (F) and nonmagnetic (N) metal layers. These multilayers can display a large relative change in the electrical resistance upon the application of an external magretic field. This effect has been widely used for commercial applications such as magnetic random access memory (MRAM), GMR read heads, and magretic position sensors [30]. 2.3.1.1 Origin of the GMR Effect GMR effect was independently discovered in 1988 in Fe/Cr/F e trilayers by a research team led by P. Griinberg [31] and in F e/Cr multilayers by the goup of A. F ert [32]. Figure 2.13 shows the GMR effect found in the Fe/Cr multilayers [32]. In these multilayers, for certain thicknesses of the Cr interlayer, the magnetizations of adjacent Fe layers are antiparallely oriented (i.e. antiferromagnetically coupled). Upon application of an external magnetic field, the resistances of the multilayers decrease drastically since the magnetizations of the two layers align in the direction of the applied field. The percent change in the resistance between the two states is defined as the GMR ratio. 27 GMR = (RAP — Rp)/Rp, (2.7) where Rpwo) is the resistance in the parallel (antiparallel) state. R/R(H=0) (Fe3run/Crl .8nm)3o (Fe3rIm/Crl .2nm)35 (Fe3nm/Cr0.9nm)4o 0. I I I I 1’ I I ) -40 -30 -20 ~10 0 10 2'0 30 40 Magnetic field (kG) Fe . . Cr Fe Figure 2.13 Magnetoresistance curves of F e/Cr multilayers at 4.2 K [32]. The GMR is related to the spin dependence of the electron conduction in ferromagnetic metals and alloys. Electron scattering depends on whether the electron spins (S,=il/2) are parallel or anti-parallel to the magnetic moment of the magnetic layers. Figure 2.14 illustrates the GMR mechanism in the two magietic states: parallel and antiparallel configurations. The resistance of the magnetic multilayers can be changed by altering the orientations of the magnetic moments of the F layers to be either parallel or antiparallel to each other. The GMR is obtained by switching the magnetic 28 moment from the antiparallel to the parallel configuration by applying the external magnetic field. Parallel Configuration Antiparallel configuration F N F W F N F "”T V Figure 2.14 Schematic representations of the scattering electrons in a magnetic multilayer. The signs + and - are for spins Sz=+1l2 and Sz=-1/2, respectively. F represents a ferromagnetic layer and N represents a non-magnetic layer, after [33]. 2.3.1.2 CIP and CPP GMR effects have been obtained in two geometries. In the first one the electric current is applied in the plane of the layer (therefore denoted as CIP), while in the second one the cm'rent flows perpendicular to the plane of the layers (therefore denoted as CPP) [34]. In general, the GMR measured in CPP is larger than that measured in CIP for the same system. The difference is due to the two different scaling lengths of the process. The scaling length of the CIP geometry is the mean free path, while the scaling length of the CPP geometry is the spin diffusion length, which is larger than the mean free path [34]. CPP-GMR is the major candidate for the next generation of read head structures since it has the advantage of development to a higher area density over CIP-GMR. 29 2.3.1.3 Exchanged-Biased Spin-Valves The exchanged-biased spin-valve structure is designed to obtain an antiparallel configuration in the multilayer [35]. In this case, the multilayer is in the form of F/N/F/AF, which is composed of two ferromagnetic layers separated by a nonmagnetic layer and an additional antiferromagretic layer (see figure 2.15). The antiferromagnetic layer acts to pin the neighboring ferromagnetic layer. When the magnetic field is increased fi’om negative to positive values, the magnetization of the free layer reverses in a very small field, while the magretization of the pinned layer remains fixed [36]. The increase of the resistance in a small field obtained in this way is now used for many low- field applications such as magnetic sensors, read heads, and Magnetic Random Access Memories. Ferromagnetic layer (free layer) Nonmagnetic layer (spacer) Ferromagnetic layer (pinned layer) Antiferromagnetic layer (pinning layer) Figure 2.15 Schematic representative of a spin-valve structure. 2.3.1.4 Current-Induced Magnetization Switching Current-induced magnetization switching (CIMS) has attracted technical interest based on the desire to switch the bits in magnetic memory by local current directly rather 30 than by the field of external wires [37]. CIMS in F/N/F metal trilayers has potential applications for writing nanopillar magnetic memory units and for CPP-MR read heads [3 8]. Since the CPP-MR is affected by spin polarization, one might expect that the switching current (15) between the parallel and antiparallel states of the F layers would be inversely proportional to the change in resistance AR. Indeed, AR was Shown to be proportional to 1/15 in Py/Cu/Py nanopillars (Py = permalloy = NimFex with x~0.2) [3 7]. 2.3.2 Correlation between GMR and Multilayer Structures In reality, the multilayer gowth mechanisms, atomic structure, and interfacial structures play important roles in the GMR effects. For example, if GMR multilayers are deposited with a rough interface or chemically intermixed layers, and/or a high concentration of crystal defects, electrons will be scattered more at the interface and/or defects, which in turn will diminish the GMR properties. A broad range of microanalysis research has been reported on the correlation between the GMR effects and the multilayer structures, including layer structures, interfacial roughness, and layer intermixing. For example, Dang and co-workers [39] studied the influence of the Co crystal structure and GMR effects in Co/Cu multilayers using nuclear magnetic resonance (NMR). Their results showed that GMR effect was enhanced by high quality layers and diminished by increased numbers of stacking faults in Co. Fullerton and co-workers [40] also showed that in Fe/Cr multilayers, the GMR was enhanced by the interfacial roughness due to strong Spin-dependent scattering of electrons. In contrast, Speriosu and co-workers [41] showed that in NiFe/Cu multilayers or spin-valves, the GMR is strongly reduced by the presence of an intermixed region at 31 the NiFe/Cu interfaces. Hence, while there is general ageement that individual layer structures will influence the MR of the multilayers [42], the roles of interface roughness, crystal defects, diffusion, and phase formation at the interface are not well understood. 2.3.3 Motivation of This Study Recently, Garcia et al. reported that the room-temperature ARs of Py/Cu/Py nanopillars were ten times larger than those of Py/Al/Py nanopillars, while the Is increased only modestly from Py/Cu/Py to Py/Al/Py [43]. These results were interpreted as suggesting a sigrificant difference in the spin-dependent scattering properties of the Py/Al interfaces and Py/Cu interfaces. To get a better understanding of the phenomenon, both Co/Al and Py/Al multilayers were studied using the CPP-MR measuring technique at MSU [38, 44]. Two parameters were characterized: (1) the enhanced specific resistance 211R} , N = (ARI, ~+ AR}, 100 (A is the area through which the CPP current flows), (2) the scattering asymmetry ymv = (ARfi/N-ARLN) /( ARI,~+AR;,N) at the F/N interface. Comparing the results of this study with the well-characterized F/N systems Py/Cu, Co/Cu, and Fe/Cr [34, 45], it was found that 2Apr/A1 is unusually large while 'YPy/A] was very small. In addition, the values of R for F/Al multilayers increased by 5- 10% after 6-11 months and then by an additional 2-7% upon annealing to 180°C [46]. Both the interfacial properties and the changes in MR might be affected by the structures and changes therein of: (1) the individual layers and interfaces, (2) interfacial 32 roughness, and/or (3) by formation of intermediate phases, as a range of intermediate phases are known to form in such systems [19]. For example, several inter-metallic phases (listed in table 2.7) are known to form in Co-Al system, as shown in the equilibrium phase diagam (figure 2.16). Vovk and co-workers [47] have found relatively rapid A19Coz and B2-ordered CoAl phase formation in Co/Al bilayer atom probe tips annealed at 300°C for 5min. Likewise, as both Fe and Ni form a range of intermetallic compounds with Al [19], one might expect additional phase formation in Py/Al multilayers. 33 00 £3.00 aaooaom 3534. 2 .8: Ease case 2.8 2a 2&5 34 c 8" o—a-nn O” G“ On” Ohm DbPPO—vpbybhbhbbFOGr bhbcbmu bbpbbbD—bptbbbbr 0°" -hIDb I b F b DID thththth _ PPF lb}. Pb bi D F} lb D F. I'D pflfht F D D.>I V. _ . n AOOHV _ u n -I ""l“l""“"l"I IIIIII“ IIIIIIII . V n m s . I g“ - fl .- n n o n A u .w n n - H In " moon m " IQI. 8.34.8. ” . . n A - Q t L . . $0.2 . n a A . ~ .I- II. a s I O8 w . . . . d I n! o \ v u \ Honda—nouns: 03254: n \ n m \ v u s . \\ I 9 A ” IHHuI... h i 1.. - S ”Ing n A u 0.82 x n r Gog“ . or. ado . . .188 I“ . r A - ‘ m OO~< . w A C v a n woe: x u \ v . . . s . n . . \s 1— n 003:“ .. \\ n u I ~ \\\ n I” I’””"“~.. \5 ‘ \ mgfifl U I u u ........ trill...F..I...I.....- ...... b...-. TIL-PI»I_IIIII->bIIPb>PI.rIIdIPIPPI>>>..-—. .b q q _ O—d c 03 cm on or on an ov on ON :38 388m Ewe; Table 2.7 Co—Al crystal structure data [19]. A19Co2 18. 1 P21/c A19Co2 monoclinic A113Co4 23.5 Cm AlBCo4 monoclinic Al3Co 25.6 N/A A15C02 28.6 P63/mmc A15C02 hexagonal AlCo ~48 to 78.5 Pm—3m Cle Cubic Furthermore, interesting MR behavior was found in Py/Cu/Py based spin-valve structures by adding a thin layer of aluminum to the center of the Spacer (Cu) layer between the two ferromagnetic layers [48]. There is a Sigrificant difference in resistance between the parallel (P) and anti-parallel (AP) states, while the difference in the resistance between the P and AP states decreased Sigiificantly by increasing the thickness of the Al layer [48]. One would expect that this behavior is attributed to the formation of an Al-Cu alloy. The Al-Cu phase diagam shown in figure 2.17 and crystal data listed in table 2.8 Show that a number of intermediate phases exist in the equilibrium or metastable states, with some of these phases still poorly defined. To understand the MR behavior variation that results from adding an Al layer to the spin-valves, both Cu/Al multilayers in different thickness ratios and selective Py/Cu/Py based spin-valve structures with Al interlayers have been studied. To understand the influence of the multilayer structures on the MR behavior and their apparent instabilities, all of the multilayer structures, including F/Al (F=Py, Co) multilayers, Cu/Al multilayers, and spin-valve structures, were characterized using cross- 35 sectional transmission electron microscopy. Individual layer and interfacial region structures were investigated by high resolution transmission electron microscopy (HRTEM) with associated fast Fourier transform (FFT). In complement, electron energy-loss spectrometry (EELS) and X-ray energy dispersive spectrometry (EXDS), performed in conjunction with scanning transmission electron microscopy (STEM), were carried out to characterize the composition nature of the multilayers. 36 .8: Same 83% 2-5 A a .N 25mm do pounce “Geckos.— 250.5. a. ... .... .... .... ... .... IIIpIhp—II.IIIII 8. av mu m» an 8&8 ...:88m Emma: 30 arnseaeduraj, 37 Table 2.8 Cu-Al crystal structure data [19]. Phase Giff/00 823:0'1 233:; Prototype “21:: 0A12Cu 31.9 to 33 .3 I4/mcm A12Cu tetragonal nlAlCu 49.8 to 52.4 Cmmm N/A orthorhombic nzAlCu 49.8 to 52.4 C2/m N/A monoclinic C1A13Cu4 55.2 to 59.8 P6/mmm N/A hexagonal C2 55.2 to 56.3 N/A N/A N/A 82 55.0 to 61.1 P63/mmc N/A hexagonal 5 59.3 to 61.9 R-3m N/A rhombohedra 71A14Cu9 62.5 to 69 P-43m A14Cu9 complex cubic BAICu3 70.6 to 82 Im-3m w 2:131)? 002 76.5 to 78 N/A N/A N/A [3’ N/A Fm-3m BiF3 bcc 38 Chapter 3 Experimental Approaches and Procedures This chapter presents the experimental approaches and procedures of all the studies throughout this thesis. The preparation procedures for the bulk Gd oxide, metallic Gd, and GdSiz reference samples for EELS quantification studies are first described. The procedures for the gowth of the silicide (including Gd and Tm silicide) nanostructures using ultra-high vacuum evaporation are then presented. The details of all the magnetic mutlilayers sputtered on Si substrates are then described. These are followed by a detailed discussion of the transmission electron microscopy (TEM) characterization, including cross-sectional TEM sample preparation, high resolution TEM (HRTEM) imaging conditions, and parameters used in image Simulation. At last, electron energy- loss spectroscopy (EELS) characterization is described by covering the type of the EELS spectrometer used in this study, spectrum collection, the curve fitting approach for quantification, Spectrum-image collection, and the simulation method used to interpret the intensity profiles from the spectrum-images. 3.1 Bulk Gd Oxide, Metallic, and Silicide Reference Samples 3.1.1 Bulk Gd203 The bulk Gd oxide TEM reference sample was prepared as follows. Micron sized Gd203 powder (fiom Alfa Aesar) was first characterized with x-ray diffraction (XRD) using a Scintag XDS 2000 diffractometer with Cu-Ka radiation to verify that the Gd203 was the Mn203 (Bixbyite) crystal structure. The powders were then gound with a mortar and pestle, and then suspended in methanol using ultrasound. Smaller sized 39 powders were picked up on a holy carbon film supported on a 200 mesh Cu TEM gid. In this way, thin platelets of Gd203 were used to carry out EELS study. 3.1.2 Bulk Metallic and Silicide Samples Produced by DC-magnetron Sputtering Bulk metallic Gd and Gd Silicides thin films were prepared using DC-magnetron triode sputtering. Si(001) wafers were cleaned using acetone, then methanol, and dried using dry N2. Alternating layers of Gd and Si were deposited on the Si substrate in a DC- magnetron triode Sputtering system with a base pressure of 3><10'8 Torr. To produce stoichiometric GdSi2 (65.4% atomic percent Si in the range of orthorhombic GdSiz, per the Gd-Si phase diagam), Si 43 nm/ Gd 75nm/ Si 85 nm/ Gd 75 nm/ Si 43nm layers were deposited. These thicknesses were monitored using a quartz crystal thickness monitor. The base Ar pressure during Sputtering was 2><10'3 Torr. After sputter deposition, the selected multilayers were annealed in a high vacuum of 5><106 torr at 900 °C for 1 hour to form Silicides. Off stoichiometric fihns did not form well defined silicide films after annealing (composed of Si and silicide gains). The annealed layers were faced to un-annealed layers to make TEM samples (section 3.4), allowing examination of both bulk Silicides and the base metal in the same TEM sample. 3.2 RE Silicide Nanostructures Grown in UHV Chamber RE silicide nanostructures were gown by evaporating the RE metals onto Si (001) substrates in an ultrahigh vacuum (UHV) chamber (base pressure 2.0X10'10torr), as illustrated in figure 3.1. Prior to deposition, the Silicon substrates were pre-cleaned by washing in acetone and isopropyl alcohol in ambient atmosphere, pre-oxidized in a UV- 40 Ozone chamber, then transferred into the vacuum chamber and repeatedly heated up to 1175 °C to remove the surface oxide. The temperature of the substrates was measured using an optical pyrometer. The cleanliness of the substrates was checked by scanning tunneling microscope (STM) imaging of the Si(001) surface, where defect densities were typically well below a few percent of a monolayer. During the deposition of the RE metals, the Si substrates were maintained at 600 °C. The RE metal was sublimed when heated by a current passing through a tungsten wire support basket and then condensed on the Si substrate. The condensation rate was calibrated by a quartz crystal thickness monitor and the metal coverage was determined by timed exposure to the source. The RE metal coverage ranged from 1 to 3 monolayers (ML) [1ML=6.78><10l4 atoms/cm2 =surface atomic density on Si(001)]. After the samples were cooled down to room temperature, they were transferred to an in-situ STM chamber for study. 600C // /' i\ W//:;‘ evaporated 8' metal Figure 3.1 Schematic representation of RE silicide nanostructure gowth experimental setup. 41 3.3 Magnetic Multilayer Synthesis All the magnetic multilayers were gown on Si(001) substrates at low temperature (from -30 to 0°C) using a dc-magnetron triode sputtering system with a base pressure of 2x 10'9 torr. Ten Co(or Py)/Al bilayers consisting of a 20 nm Co (or Py) layer and a 10 nm Al layer were sputtered on Si(100) substrates. For comparison with the as-sputtered samples, some multilayers were annealed for less than 5 min at 180°C after sputtering in the same chamber. Alternating Cu/Al multilayers were gown in the same dc-magnetron triode Sputtering system with similar conditions. These multilayers were sputtered in different thickness ratios: (1) Cu(8nm)/Al(10nm) with an overall composition of 52.9at. % Cu, (2) Cu(5nm)/Al(3nm) with an overall composition of 70.1at. % Cu. Some of the Cu(8nm)/Al(10nm) multilayers were annealed for less than 5 min at 180°C for comparison with samples in the as-sputtered condition. The two spin valves studied had the following configurations, where all the values are given in nanometers: Nb(1 50)Cu(5)FeMn(8)Py(24)Cu(l 0)Al(x)Cu(l 0)Py(24)Cu(5)Nb(50), where x=10 for the as-sputtered sample and x=30 for the annealed sample. 3.4 TEM Characterization Most of the Specimens in this study were characterized using cross-sectional TEM. The cross-sectional TEM sample preparation is introduced in this section. This is followed by a description of HRTEM imaging conditions and image simulation analysis. 42 3.4.1 Cross-sectional TEM Sample Preparation TEM cross-sectional samples were prepared using the sandwich technique by joining the thin films on the Si substrates (including bulk thin films, silicide nanostructures, and magnetic multilayers) face-to-face and sectioning along a Si<110> direction lying in the surface plane of the Si substrate. The samples of the sputtered thin films and the silicide nanostructures were separately joined using G1 epoxy (Gatan), which can be cured at 100 °C for 5 min. The samples of the magnetic multilayers were joined using M-bondTM 610 epoxy (SP1) and cured at room temperature for a relatively long time (i.e. 24h) to prevent the samples from experiencing high temperature during sample preparation. The cross sections were glued onto molybdenum (Mo) rings, pre- thinned by mechanical thinning and dirnpling, and then low angle ion milled to electron transparency. Figure 3.2 shows a schematic representation of the entire process of the sample preparation. The details of the procedures to form cross-sectional TEM samples are as follows [49]: 1. Section thin film sample on substrate into about 4 mm X 7 mm slabs with a diamond scribing pen. 2. Clean slabs in acetone, followed by a rinse with ethanol. 3. Glue two slabs of the thin film face-to-face with G1 epoxy. The epoxy was cured by heating to 100 °C for 5 min on a hot plate. (The magretic multilayers were joined using M-bondTM 610 epoxy and cured at room temperature for a relatively long time i.e 24 h.) 4. Slice the slab perpendicular to the gowth plane into 800~1000 um thin slabs using a diamond saw (Struers Accuton-5 operated at 3000-rpm wheel speed with 43 a 0.05 mm/sec feed rate). 5. Glue Slices onto Mo rings with G1 epoxy and cure by heating to 100 °C for 5 min again. (The cross sections of the multilayer samples were bonded to M0 rings using M-bondTM 610 epoxy and cured at room temperature for a relatively long time i.e 24 h.) 6. Mount the Slab and Mo ring to the center of a sapphire flat with a thin layer of CrystalbondTM mounting wax. The CrystalbondTM wax is heated to 130 °C on a hot plate for 5 min. The melted wax is transferred using a toothpick onto the backside of the Mo ring (which is briefly held on a glass slide on the edge of the hot plate. Temperature measurements Show this area to be at ~ 80 °C.). A sapphire flat, preheated on the edge of the hot plate (~ 80 °C), is placed onto the ring and then removed with the sample and ring. This entire bonding process takes place less than 30 seconds. The crystal bond solidifies in a matter of seconds. 7. Grind and polish the open side of the sample assembly using a series of (15, 6, 3, l um) diamond lapping films (Allied) until the slice thickness is around 100 um. 8. Attach (using a screw fitting) and center the sample assembly on a VCRTM Group D500i Dimpler platen using an eccentric platen assembly. A 15.5 mm stainless steel dimpling wheel and 1 urn diamond Slurry are used. The dimpling is done until a center thickness of about 20 pm is reached. 9. Remove the sample assembly from sapphire flat with acetone to dissolve the crystal bond. 10. Rinse the sample assembly in ethanol and air-dry. ll. Ion mill using a Gatan M691 PIPSTM ion mill with a Gatan DuoPostTM support. The center of sandwich interface is ion milled at fl°~ :l:4° beam incident angle with 4 keV Ar+ ions until perforation. / glue cut 1n half face-to—face cross-sectional view mount on Ar‘ gun dimple afier grinding Mo ring I, Low angle ion milling {-— (——— cross-sectional TEM sample Figure 3.2 Schematic representation of the process of preparing a cross-sectional TEM sample. 3.4.2 High Resolution TEM HRTEM was carried out using a JEOL JEM2200FS field emission gun TEM/STEM working at 200KV, which has a C5, of 0.5 mm, and hence, a theoretical point-to-point Scherzer resolution of 0.19 nm. During observation, Si substrates were tilted to Si<110> zones to orient the thin film (either bulk Silicides, silicide nanostructures, or multilayers) edge-on. HRTEM images were collected using the Gatan DigitalMicrogaphTM microscope control and imaging software. Fast Fourier transform 45 (F FT) diffractogams were obtained and analyzed using ImageJTM. By comparison with the diffraction pattern simulated using J EMS (developed by Pierre A. Stadelmann at Center Interdisciplinaire de Microscopic Electronique in Switzerland [50]), crystal zones and structures were identified. Some of the structural features were too small to obtain good quality F FTs. In this case, HRTEM image simulation using I EMS was performed to interpret the crystal structures. The selected microscope parameters are as follows: accelerating voltage 200 keV, Spherical aberration Cs 0.5 mm, chromatic aberration Cc 1.0m, with thickness and defocus as variables. 3.5 EELS Characterization 3.5.1 Design of Electron Energy-Loss Spectrometers Two types of EELS spectrometers are presently used in TEM: the omega (9) filter (LEO) and the magnetic prism spectrometer (Gatan). The {2 filter disperses the electrons in the TEM column. It is placed between the intermediate and projector lenses, as Shown in figure 3.3a [2]. The 0 filter is used mainly for energy-filtered imaging although spectra can be obtained. It is a specialized spectrometer that has to be built into the microscope column rather than an added option. The Gatan Image Filter (GIF) (shown in figure 3.3b) is mounted under the microscope column without changing the optical path in the microscope [2]. The JEOL JEM2200FS transmission electron microscope in the Center for Advanced Microscopy at MSU is equipped with an in- column Q energy filter allowing EELS and energy-filtered TEM (EFTEM). 46 f To projector lens crossover object plane of spectrometer .... 8L rrj‘)‘ Sextupoles Figure 3.3 Schematic overview of the in-column (2 filter (a) and post-column Gatan Image Filter (b) after [2]. 3.5.2 EELS Spectrum Collection Electron energy-loss spectra were collected in STEM mode with a ~0.5nm size probe. The zero-loss peak was collected before each analysis to confirm the spectrometer had good energy resolution (between 1.0 to 1.5 eV FWHM, see figure 3.4). This was done with a minimum exposure time (i.e. 0.01s) to avoid damaging the CCD camera. The Gd M45-edge was selected for quantification analysis to understand the electronic structures of Gd in the samples. Since it is a high energy-loss edge (~1185eV), this edge requires a 10 sec acquisition time to achieve suitable counting statistics. 47 Intensity (arb. units) . I ‘ ‘ 5 I I ' l ‘ . l i . i i u..- :L i - a 4 . . = 9 . I : - .' ‘ i __ l 2 I 17 l l I 1 20 40 60 80 100 120 I40 100 180 Figure 3.4 Zero-loss peak with a full width atelilalf maximum (FWHM) ~1.0 eV shows that the EELS spectrometer is well-aligned. 3.5.3 Gd M-Edge EELS Spectrum Curve Fitting For quantitative analysis of the Gd M45-edge EELS spectrum, spectrum model fitting techniques developed by Manoubi et al. [51] were used to extract useful information from experimental spectra. The experimental spectra of M45-edges were fitted as a superposition of a pre- edge power-law backgound, with a set of Lorentzian profiles to account for the white- lines, and a continutun term corresponding to the transitions to empty states in the continuum modeled by a Hartree-Slater function of cross sections extracted from Gatan DigitalMicrogaph [52] (figure 3.5). 3.5.4 Spectrum Image Collection To analyze the interfacial chemical nature of the F/Al multilayers, a series of Spatially collected EELS spectra, referred to as a EELS Spectrum image, were collected by scanning an ~05 nm diameter (FWHM) probe across the interfaces of the F/N 48 multilayers. Each spectrum was composed of 1024 EELS channels of 0.11 eV width. Low loss spectrum images with plasmon peaks were acquired over scans of ~50nm across F/N/F/N/F layers. Core loss spectrum images with white lines for Co (or Ni, Fe) were acquired using scans of ~20nm across the F/N/F layers. During the acquisition process, specimen drift caused by thermal effects was corrected using Gatan DigitalMicrogaphTM by periodically scanning over a cross-correlated reference region in the survey images. Intensity (arb. units) 1 I'l r GdSi2 EELS spectrum curve fitting M 5. l' \—~~ I l I I l 1 I l 160 l 180 1200 1220 1240 Energy loss (eV) Figure 3.5 A typical EELS spectrum of M45 edge (black line) from GdSi; fitted as a superposition of a power-law background (straight line), cross sections from Hartree- Slater model (rising line), and a set of Lorentzian profiles (gay lines). The dotted curve shows a good curve fitting result. 3.5.5 Intensity Profiles and Simulation 49 The acquired spectrum images were processed in the Spectrum Imaging package in Gatan DigitalMicrogaphTM. Each spectrum in the image (core loss spectrum image) was extracted and then normalized by subtracting a pre-edge power-law backgound. The intensities of the peaks (plasmon peaks and core loss peaks after backgound subtraction) were then plotted as a function of distance. The intensity profiles of the as- sputtered and annealed F /Al multilayers were then compared. To facilitate an understanding of the experimental F L-edge intensity profiles, theoretical intensity profiles across perfect F/Al/F interfaces (figure 3.6) were simulated using an electron beam with a 0.5 nm (FWHM) Gaussian profile. Normalized intensity rjmvfirf'vvv' v 'u v' 2.5 5 7.5 10 I25 Distance (nm) Figure 3.6 Simulation of an F (Co, Ni, or Fe) L-edge intensity profile with a 0.5nm FWHM of Gaussian electron beam profile across idealized F/Al interface. 50 Chapter 4 Results from RE Silicide N anostructure Studies In this chapter, the results of the HRTEM and EELS studies of Gd and Tm silicide nanostructures are presented. The crystal structure study of the bulk phases, including Gd oxide, metallic Gd and GdSlz in thin film form, are presented first. This is followed by cross-sectional HRTEM study of Gd and Tm silicide nanostructures. EELS quantification study of the Gd silicide nanostructures compared to the bulk phases will be presented last. 4.1 Gd Oxide, Bulk Metallic and Silicide Crystal Structure Study To understand the extent that the nanostructures vary from the bulk materials, bulk samples, including Gd oxide, metallic Gd, and GdSi; in thin film form, were used as references for crystal and electronic structure study. The crystal structure of the as-received Gd203 powders was studied by 0-20 x-ray diffraction (XRD) scan. The scan is presented in figure 4.1, which indicates that the Gd203 has a lattice constant a=10.813 A, space goup Ia3, and is consistent with the prototype of Mn203 (Bixbyite) structure [50]. Figure 4.23 Shows three gound GdzO3 powder particles on a carbon holy film. These particles are in the range of a hundred nanometers in diameter. Insets (b) and (c) Show that the individual particles can be either single crystals or polycrystals. The thin areas of the Gd203 particles are suitable for EELS study. 51 60 50 - Counts (CPS) 20 25 30 35 40 45 50 55 60 65 70 2 0 (degree) Figure 4.1 X-ray diffraction Shows the bcc Gd203 agee with the prototype of Mn203 (Bixbyite) structure. — 100 nm Figure 4.2 (a) Phase contrast TEM image of Gd203 powder particles on a holy carbon film. Insets (b) and (c) Show HRTEM images of a Single gain and poly- gain atomic lattice from two powder particles respectively. 52 Figure 4.3 shows a cross-sectional TEM image (a) and a corresponding selected area electron diffraction (SAD) pattern (b) of the sputtered Si/Gd layers before annealing. The Si is amorphous and the Gd displays fine crystalline gains. The selected area electron diffraction pattern shows polycrystalline diffraction rings corresponding to the hexagonal crystal structure of metallic Gd (with c/a ratio of 1.59). It is known that this RE metal is reactive and oxidizes readily. Nevertheless, subsequent EELS examination of this layer revealed no obvious oxygen K edge. As a result, metallic Gd EELS spectra can be collected from the Gd layers. 102 100 Amorphous 110 Si . Figure 4.3 (a) Phase contrast TEM image and (b) corresponding selected area electron diffraction pattern of amorphous Si and metallic Gd layers alternately sputtered on a Si(001) substrate. Upon annealing at 900 °C for 1 hour in vacuum, the Gd/ Si layers transformed to GdSi; by inter-diffusion of Si and Gd. Figure 4.4a presents a low magnification phase contrast TEM image of the post-deposition annealed bulk GdSi; layer formed on the Si(001) substrate. The GdSi; layer ranged in thickness from 100 to 200 nm and formed as a relatively large gained structure, with individual gains traversing the entire layer. A SAD pattern from one gain of the bulk GdSi; layer shows a single crystal [100] zone (seen figure 4.4b). The HRTEM image in figure 4.4c shows the GdSi; layer imaged 53 along this [100] zone axis. The imaged atomic lattice is consistent with the projection viewed along the orthorhombic GdSi2[100], with a half-atom shift between every two horizontal rows of atoms (aabb stacking in figure 2.11). It has been reported that both the hexagonal GdSi” and the orthorhombic GdSi; can be formed by annealing Gd thin films on Si substrate, depending on the thickness of Gd layer and the Si orientation (Si(001) and Si(111)) [53, 54]. In our case, only the orthorhombic phase GdSi; was found to exist, which indicates that the orthorhombic phase is energetically favorable under these conditions. This bulk GdSi; will be used as a reference for comparison with the Gd silicide nanostructures. 54 .7 4“" .35»! , . .5055” ~ .9” Bulk GdSi2 layer Si Substrate 55 Figure 4.4 (a) Phase contrast image of post-annealed bulk GdSiz layer formed on the Si(001) substrate. (b) Selected area diffraction pattern from one gain of the layer with arrows presenting strong diffraction spots from the orthorhombic phase of GdSiz. (c) HRTEM image showing the [100] orthorhombic zone axis from the GdSiz. The atomic lattice symmetry is indicted by the stacking of (001) planes in an aabb manner. 4.2 HRTEM Studies of Gd and Tm Silicide Nanostructures This section will present results from cross-sectional HRTEM study of both Gd and Tm silicide nanostructures. To correlate the silicide phase crystal structures with their morphologies, some STM topogaphic images are presented before showing the HRTEM results. The Gd silicide nanostructures imaged are from one TEM sample, which means all the nanostructures were gown with the same metal coverage and heating process. In contrast, the Tm silicide nanostructures imaged were from different samples with different gowth conditions, which are described separately. Cross- sectional HRTEM images, canied out in conjunction with FFT, are assessed with image simulations to interpret the experimental images and identify the crystal structures. 4.2.1 Gd Silicide Nanostructures The Gd silicide nanostructures imaged were from one TEM thin foil. The nanostructures were gown on a substrate heated to 650°C with a metal coverage of 1.5ML. This was followed by post-deposition annealing for 20 min at the gowth temperature. An STM image of this specific sample rendered in 3-D (figure 4.5) shows a large number of islands, mostly compact, with a small fraction of high aspect ratio nanowires. The line profile shows the dimension of an island with varying thickness up to 3nm from the bottom of adjacent trenches (these trenches are due to local Silicon depletion during the formation of the silicide). 56 0)) “height (nm) distance (nm) Figure 4.5 (a) 3D topogaphy STM image of GdSi2 nanostructures gown on Si(001) with (b) a cross-sectional line profile, indicated by the line on (a). The cross-sectional phase contrast image in figure 4.6a shows the entirety of a GdSi; nanostructure, approximately 70 nm in width and 3-4 nm in height. The atomic resolution HRTEM image in figure 4.6b exhibits atomic planes with aabb stacking, which is consistent with the orth/tet phase of GdSi2 viewed along the [100] zone axis (shown in figure 2.11b). An FFI‘, which includes information from both the GdSi: and Si (figure 4.6c), shows that there is an angle of approximately 2° between the GdSi2(020) and the Si(220), suggesting that the epitaxial relationship deviates slightly from a perfect GdSiz(001) // Si(001) and GdSi2[100]// Si[110]. This tilt most likely reflects a strain accommodation mechanism. There is a lattice mismatch of approximately 5% between the GdSi2(020) and Si(220) measured by the FFT. The lattice parameters measured from this nanostructure are a (and/or b) = 0.40 nm, c = 1.35 nm, which agee well with the expected values for bulk silicide lattice parameters within experimental error (see table 4.1). 57 - 8 >4 -.--. .P . ‘1} iv I'K'.'” Ia.‘ V";’.A{,I‘V\"‘ ""4"“.f’ "hr ' ‘ In ~ . a" If '31?» ' ' 13:1.”"93‘ .5:- I (idSi2(020)i-I t . All " suzztijf ‘ffsiroogi ' lSi2(00-l)/7' ‘ Nfifi‘OStructure ._'.-‘1Ir-- Si substrate ' l _. Figure 4.6(a) Cross sectional phase contrast TEM image of a GdSiz nanostructure gown on Si(001). (b) Higher magrification HRTEM image showing the interface between the GdSiz and Si. (c) FFT of area shown in (b). Table 4.1 Lattice parameters of the bulk GdSi2 phases and those measured from the three imaged nanostructures. (I, II, III refers to the imaged nanostructures) Structure a(nm) b(nm) c(nm) Mismatch with Si (001) a(%) b(%) C(%) Bulk Hex 0.388 0.417 1.0 8.6 phases Tetra 0.410 1.361 6.77 6.77 OIth 0.409 0.401 1.344 6.51 4.43 Measured Hex 0.39III 0.42 1.6 8.0 Tetra/orth 0.40I 1.35 5 0.41II 1.36 6.8 0.39m 1.34 Figure 4.7a shows a cross-sectional phase contrast TEM image of another GdSi; nanostructure, approximately 70 nm in width with varying thicknesses up to 7 nm. The 58 atomic resolution HRTEM image in figure 4.7b again exhibits an aabb stacking, which is consistent with the orth/tet structure viewed along the [100] zone axis (shown in figure 2.11b). An FFT diffractogam in figure 4.7c shows the Si[110] zone and the orth/tet phase of GdSi2[100] zone. Similarly, the epitaxial relationship identified deviates slightly from a perfect GdSi2(001) // Si(001) and GdSi2[100]// Si[110], with a tilting angle of ~2° between GdSi2(020) and Si(220). The lattice mismatch parallel to the interface measured by the FFT is approximately 6.8%. The lattice parameters measured from this nanostructure are a (and/or b) = 0.41 nm, e = 1.36 nm, which agee with the expected values for bulk silicide lattice parameters within experimental error (see table 4.1). ' " Nanostructure . T l-(lnin' Si (220) I . /fl ,cdsrzmzo t I I 1 (.(lSIfllfi-I) Si 002 Figure 4.7 (a) Cross sectional phase contrast TEM image of a GdSi; nanostructure with multiple thicknesses. (b) Higher magnification HRTEM image shows an aabb lattice stacking, consistent with the orth/tet phase of GdSi; viewed along the [100] zone axis. (c) F FT from area (b) confirms the crystal structure of the orth/tet phase of GdSiz. 59 In both cases, the fact that these relatively thick islands are entirely orthorhombic phase of GdSi; is consistent with previous work [20]. Since the sample experienced post- deposition annealing, this observation suggests that the orthorhombic phase of GdSi; is the energetically favorable silicide phase. Figure 4.8a shows an HRTEM image of a third GdSi; nanostructure, approximately 42 nm in width with a wavy top with thickness up to 10m. The structure is tilted relative to the Si(001) surface. The structure contains two types of atomic stacking, separated near the dashed horizontal line Shown at the left hand side of the image. The top part of the structure exhibits an aabb atomic stacking. The FFT from area 1 (in inset) Shows a tilted orth/tet GdSi2[100] zone. The bottom part shows a different atomic stacking with rectangular symmetry. This stacking is consistent with the hexagonal phase of GdSiz.x viewed along the [100] zone (shown in figure 2113). An FFT diffractogam from area 2 (in inset) shows a hexagonal GdSiz.x [100] zone. The FFT also shows that there is an approximately 4° angle between the GdSi2-x(002) and Si(220), which suggests that the epitaxial relationship deviates from hexagonal GdSi2-x(OlO) // Si(001) and GdSi2-,.[100]// Si[110]. There is a lattice mismatch of approximately 8% between the hexagonal GdSi2-x(002) and Si(220), measured by the FFT fi’om area 2. The lattice parameters measured from the hexagonal phase are a = 0.39 nm and c = 0.42 nm, and the lattice parameters measured from orth/tet phase are a (and/or b) = 0.39 nm and c = 1.34 nm, both of which agee with the expected value for bulk silicide lattice parameters within experimental error (see table 4.1). The relatively large mismatch (8%) between the hexagonal GdSi2(002) and Si(220) requires a large strain accommodation at the interface. A periodic array of strain 60 contrast features along the interface (marked by arrows) is clearly resolved in figure 4.8b (enlarged from are 3 in figure 4.8a), which is consistent for an array of misfit dislocations. The spacing of the strain contrast (a pair of dark and bright constrast features) is approximately 2 nm (i.e. 5-6 Si(l 10) planes). This spacing is very close to the spacing between the thinner nanowires that gow in bundles at lower metal coverages (which is reported in the range 5-7 unit cells, i.e. 2-3 nm) [55]. The Burger circuit shows that there are approximately 9 GdSi2-x(001) planes on every 10 Si(l 10) planes, resulting in a semi- coherent interface. The Burger vector b measured is 1/2<110> in Si, existing in every two contrast features. However, the 9:10 mismatch, with the observation of approximately 5-6 Si(110) planes suggests that the observed dislocations are partial dislocations. It should be noted that if these are in fact partials, they would have b=l/4<1 11> in the Si, with displacements out of the interface plane. Based on the measured lattice mismatch between Si and Silicides, the spacing between the 1/2<110> dislocations should be 4.48nm which covers two pairs of dark and bright contrast periods. 61 ,‘- ." I A.--.',Nanostruct'ure.o ' ' _ . (“5‘2 A (020) 84002), i P05521016) . . . 2 ‘\‘ ’/ I I I I I I I ,l .I I Figure 4.8 (a) An HRTEM image of a GdSi; nanostructure with a wavy top. Two different lattice symmetries are observed in the nanostructure on the two sides of the dashed horizontal line. Above this line an aabb stacking is displayed, and an FFT fi'om area 1 shows a tilted orthorhombic GdSi2[100] zone. Below the line, close to the Si substrate a rectangular symmetry with aaaa stacking is observed. An FFT from area 2 shows a hexagonal GdSi2_x[100] zone. (b) An enlarged image from area 3 illustrates the contrast from mismatch dislocations at the interface, as well as a periodic array of strain contrast marked by arrows. 62 According to the analysis from these GdSi; nanostructures, both hexagonal GdSiz- x and orth/tet GdSiz can self-assemble on Si(lOO). The F FT analysis illustrates that the lattice parameters of the nanostructure phases do not deviate significantly from the bulk phases. The strain caused by the lattice mismatch between the Silicides and Si substrates is accommodated by tilting of the orth/tet phase or the introducing dislocations between the Si and the hexagonal phase. The fact that two phases coexist in the same silicide island is not surprising, as similar behavior has been seen for Dy silicide islands [56]. The fact of one phase overgows another suggests a strain relaxation mechanism is acting in these nanostructures. We can define a two dimensional (2D) mismatch as the product of the mismatches in the orthogonal directions. Taking the bulk data from Table 4.1 for the hex phase: 1.01 x 1.086 = 1.097 -) 9.7% for 2D mismatch. In a Similar manner, the 2D mismatch between the substrate and the orthorhombic phase is 11%, and the 2D mismatch between the two Silicides is 1.3%. It can be seen that the mismatch between the hexagonal and orthorhombic phases is very small, and Should result in a coherent or semi-coherent interface. A planar, semi-coherent interface is likely lower in energy than a more incoherent interface associated with tilting, which will require a combination of dislocation and/or atomic ledges to accommodate the mismatch. As the hexagonal phase has a smaller 2-D lattice mismatch with the Si substrate than the orthorhombic phase, having the hexagonal phase act as a buffer layer at the interface between the orthorhombic phase and the substrate may avoid complex interfaces and provide strain relief. 63 Another possible explanation for the observed overgowth of the orthorhombic phase on hexagonal may be that it results from the intersection of two gowing structures, since other work has Shown that the orthorhombic phase can nucleate at the intersections of intersecting hexagonal islands [20]. Thus, it is possible that this cross section reveals the overgowth of the orthorhombic phase over a hexagonal island. 4.2.2 Tm Silicide Nanostructures 4.2.2.1 Nanostructure Morphology The Tm silicide nanostructures were gown at approximately 600°C with a metal coverage of l to 3 ML, either with or without post-deposition annealing. Figure 4.9a shows an STM image of nanostructures gown with 2.8 ML metal coverage without post- deposition annealing (Sample Tm-A). There are two basic morphologies of nanostructures gown on the Si surface: nanowires and islands. The nanowires are ~ 10 nm wide and 0.3 to 1.3 nm high. The islands typically have much lower aspect ratios and are often observed at nanowire intersections. It is difficult to discriminate islands and nanowires on the basis of height, but the island widths are typically 20 nm or more. Figure 1b shows the cross sectional profile along the line displayed in the image. Three nanostructures are seen: (1) a multiple layer thick nanowire, (2) a one layer high nanowire ~0.34 nrn high and 10 nm wide, and (3) an island ~30 nm wide. Typical features exhibited by islands include a minimum height of 1 nm, a curved top, and adjacent trenches that indicate local depletion of Si. Some of the islands Show a periodic modulation in height. One of these islands, from a sample gown at 1 ML metal coverage with a post-deposition annealing at 600°C for 30 min, is shown in figure 4.10 (Sample Tm-B). This island shows a modulation of about 0.1 nm in amplitude and a lateral period of about 10nm. This modulation period in diflerent islands ranges from 7 to 15 nm. The modulation is seen in both filled and empty states STM images, strongly suggesting that it is structural rather than electronic in origin. (a) 10.5 nm (2) Figure 4.9 (a) An STM image of Tm silicide nanostructures on Si(001) with 2.8 ML Tm coverage. (b) A cross sectional profile along the line shown in (a). Figure 4.10 (a) An STM image of Tm silicide nanoisland with a modulation in height. (b) Cross-sectional profile along the line shown in (a). A third sample (Tm-C) was gown under very similar conditions to Tm-B: 1 ML Tm deposited at 600°C with 20 minutes post annealing. However, this sample was also capped with a layer of amorphous Can in an attempt to decrease the amount of post gowth oxidation that occurs once the samples were removed from the UHV system. The most sigrificant difference between samples Tm-A and Tm-B was that Tm-A was gown to a higher metal coverage and consequently had a tendency to Show a higher fi'action of silicide islands in the STM images. 4.2.2.2 HRTEM Image Simulation HRTEM image simulations were carried out to interpret the stacking symmetries of the projections along various axes of the potential Tm silicide phases. These the simulated images were then compared with the experimental HRTEM images so that the crystal structures could be identified. Figure 4.11 shows a comparison of the projections and simulated images of the hexagonal Tm3Si5 and orthorhombic TmSi along different axes. If either of these two structures forms nanowires, figures 4.11 a and c would be expected to be the side view of the nanowires, while figures 4.11 b and (1 would be expected to be the “end on” views, based on the anisotropic gowth mismatch with these phases. The simulated image along the hexagonal Tm3Si5[001] (figure 4.113) displays a six-fold symmetry, while the simulated image along the orthorhombic TmSi[lOO] (figure 4.11c) shows an arrangement of flattened and distorted hexagons, and also a clear band in dark contrast between every two atomic layers (note that the high Z RE cations dominate the contrast). The simulated image along the hexagonal Tm3Si5[100] (figure 4.1 lb) Shows a rectangular symmetry with aaaa stacking, while the simulated image along the orthorhombic TmSi[OOl] (figure 4.11d) shows a aabb stacking. These differences facilitate discrimination of the two phases in HRTEM cross-sectional images. 66 3 '3 '3 . . 2 . 2 . . . 0 . 131919911011 along Image simulation Tm3Sls [001] along TmJSis [001] Projection along Image simulation TmSi [100] along TmSi [100] Projection along Image simulation Tm3Si5 [100] along T1113815 [100] Projection along Image simulation TmSi [001] along TmSi [001] Figure 4.11 A comparison of the projections and corresponding simulated lattice images of the hexagonal Tm3$i5 (a and b) and orthorhombic TmSi (c and d) imaged down different axes. 4.2.2.3 Results from HRTEM Imaging Figure 4.12a shows a cross-sectional HRTEM image of a Tm silicide nanostructure from sample Tm-B, approximately 10nm in width and less than 1 nm in height. The enlarged lattice image Shows a rectangular stacking symmetry with only two atomic layers clearly resolved. This makes it impossible to determine the crystal structure since it does not have a complete unit cell: the rectangular symmetry with only two cation layers appears in both figures 4.11 b and d. The corresponding FFT (figure 4.12c) shows streaked diffraction peaks since very limited periodicities are present. 67 Figure 4.12 (a) HRTEM image shows a Tm silicide nanostructure approximately 10 nm in width and less than 1 nm in height. (b) Enlarged lattice image shows a rectangular stacking symmetry. (c) FFT shows streaked diffraction peaks. Figure 4.13a shows a cross—sectional HRTEM image from sample Tm-A of a Tm silicide nanostructure approximately 35 nm in width and approximately 1 nm in height. Two sections of the nanostructure are noted. To the left, a section approximately 6 nm in width is observed. This section is separated by approximately 4 nm of Si from a much broader section approximately 26 nm in width. This nanostructure has similar dimensions to the island seen in the STM image in figure 4.10. The rectangular symmetry of the lattice stacking (aaaa stacking) seen in the enlarged atomic lattice image (figure 4.11b) and the corresponding FFT (figure 4.130) are consistent with those of the hexagonal Tm3Si5 viewed along [100] (i.e. figure 4.11b). FFT analysis reveals that the epitaxial relationship is hexagonal Tm3Si5(100) // Si(001) and Tm38i5[001] // Si[110] with an angular deviation as much as 5°. Because the nanostructure is so thin, only lateral periodicities in one direction result in good diffraction peaks, and only the lattice 68 parameter c, measured to be 0.40 nm, can be determined. This value is consistent with the bulk hexagonal silicide value 0.407 nm (listed in table 4.2). Figure 4.13d shows an x- ray energy dispersive spectrum that confirms the nanostructure contains significant Tm. Figure 4.14a shows a cross-sectional HRTEM image from sample Tm-C of an elongated Tm silicide nanostructure approximately 90 nm wide and up to 3nm thick, with a Can capping layer. The enlarged lattice image in figure 4.14b shows stacking is similar to the simulated image of the orthorhombic TmSi viewed along [100]. The corresponding FFT (figure 4.14c) shows a TmSi [100] zone. The lattice parameter c is measured to be 0.38nm, which is consistent with the bulk value 0.378nm (listed in table 4.2). To get clearer contrast in the lattice image, the FFT was filtered by selecting the TmSi and Si diffraction peaks respectively and filtered images were obtained from the nanostructure (figure 4.14d) and Si (figure 4.14e). Figure 4.14d shows the stacking has a roughly six-fold symmetry, but more significantly, a dark contrast band between every two atomic layers, which is consistent with the simulated image (inset) viewed along the orthorhombic TmSi viewed along [100]. Thus, this nanostructure can be interpreted as a nanowire with the orthorhombic TmSi crystal structure viewed on side. 69 ‘- .4 (c) A .' Sl‘lii38i5(ltm,‘/Si(tl(|2) ' < ' t 5 l . ‘ . I , /‘. s ‘ f _'Ttii35i5(001) -‘ “(22“, Si substrate C "5 _ Si =3 3 3 = 5 - O U 0 Tm Tm Tm 1 l l A l l J _L [y MA 1““ l 0 1 2 3 4 5 6 7 8 9 10 Figure 4.13 (a) Cross sectional HRTEM image of a Tm silicide nanostructure on Si(001). (b) Enlarged atomic lattice shows rectangular symmetry. (c) Corresponding FFT from area (b). ((1) X-ray energy dispersive spectrum shows chemical nature of the nanostructure. 70 Table 4.2 Lattice parameters of the bulk Tm silicide phases and those measured from the imaged nanostructures. Structure a (nm) b (mm) 0 (11m) Mismatch with Si (001) a(%) b(%) c (%) Bulk Orth 0.418 1.035 0.378 +8.9 +170 -l.6 phases Hex 0.377 0.407 -l.9 +6.0 Orth 0.40 1.32 0.38 +4.2 -1.0 Measured Hex 0.40 +4.2 CaF7 VIA-Nanostructure _ (10' nm .‘ TmSii {LU} ‘/ V CLO-‘4’vlfmno-0t-04s - 0 Figure 4.14(a) Cross-sectional HRTEM image of Tm silicide nanostructure from an elongated Tm silicide nanostructure ~90 nm wide and up to 3nm thick between Si substrate and Can layer. (b) Enlarged lattice image shows a stacking similar with the simulated image of orthorhombic TmSi viewed along [100]. (c) Corresponding FFT shows a TmSi [100] zone. ((1) and (6) shows a comparison of FF T filtered lattice image from the nanostructure and Si, which confirms the stacking in the nanostructure is consistent with the simulated image of orthorhombic TmSi viewed along [100] (inset in (d)). 71 Figure 4.15a shows an HRTEM image of another Tm silicide nanostructure from sample Tm-C, which is approximately 30 nm in width and approximately 2 nm in height. The lattice in the left in the image (area 1) shows a stacking symmetry with a half-atom horizontal shift between every two rows of atoms, which is consistent with the schematic projection along the orthorhombic TmSi[OOI] in figure 4.11d. The corresponding FFT (figure 4.15b) is also consistent with the orthorhombic TmSi[OOI] zone pattern. Since the Tm silicide lattice is limited in height, the diffraction peaks for TmSi {040} are streaked in the FFT. It is noted that the lattice in the right of the image (area 2) shows different apparent symmetry. However, the FF T (figure 4.15c) reveals a tilted orthorhombic TmSi[OOI] zone, which indicates the structure is still orthorhombic but slightly tilted. This suggests some strain accommodation is taking place in the nanostructure and the epitaxy on Si(001) is not ideal. The lattice parameters measured fiom the FFT are a=0.40 nm, b=1.32 nm, which vary from the bulk values a=0.418 nm, b=1.035 nm to some extent (listed in table 4.2). As a result, the large lattice mismatch expected with the orthorhombic a axis (8.9%) turn out to be ~4.2%, which may suggest a lattice distortion occurred. This magnitude of lattice mismatch may not constrain the lateral growth as expected from the bulk orthorhombic phase, which may result in a relatively broader lateral dimension. Based on the analysis from figure 4.14, orthorhombic TmSi may be the phase that forms nanowires on Si. As a result, this nanostructure in figure 4.15 may be a nanowire viewed end on. The relatively small lattice mismatch (4.2%) may not significantly confine the lateral growth, which results in a width of approximately 30 nm for this nanowire. 72 'll 1.. pppppp ' :‘,"-:\‘1t::1)l +\~. ' ' ..., ,. _ limit—111M". ‘ " U , . v I1 a I " VT\ I n ‘ . . lmMHHm wool) ...... , \\ ‘ " ,. v ’ lmxn‘lllm / Figure 4.15 (a) Cross—sectional HRTEM image of Tm silicide nanostructure on Si(001) shows a stacking symmetry with a half-atom shift between every two rows of atoms consistent with the orthorhombic TmSi [001] axis. (b) and (c) are the corresponding FFTS from areas (1) and (2). A significant portion (2 out of 3) of nanostructures (in all the three samples) were observed to contain more than one type of stacking lattice, which may suggest they are made up of different crystal structures. Figure 4.16 shows two examples of this type of nanostructure, which are from sample Tm—C. Figure 4163 shows an HRTEM image of a Tm silicide nanostructure approximately 25 nm wide and approximately 2nm thick. Two different lattice stackings are observed. The bottom two layers of the structure show stacking consistent with the simulated image of the orthorhombic TmSi viewed along [100], while the upper stacking is consistent with simulated image of hexagonal Tm3Si5 viewed along [100]. Figure 4.16b shows an HRTEM image of a Tm silicide nanostructure approximately 10 nm Wide. Again, the bottom part of the nanostructure has stacking consistent with the simulated image of the orthorhombic TmSi viewed along [100], while the upper two layers are consistent with simulated image of hexagonal Tm3Si5 viewed along [100]. FFTS cannot be used to determine the crystal structures due 73 to the limited periodicities. The poor contrast between the different stacking may suggest these nanostructures exhibit two different crystal structures. C a F2 Nanostructure Si substrate 5 nm €an _ , ._ .7 ,_ , ./ Nanostructure ,_ 195032-“-‘*““W’-"* J?" ' ' ‘ ‘ ‘.<"“>‘\\\\\s¢\s{\\.\‘"i:\mt. Si substrate?) - “ ‘ ' Stun 31... Figure 4.16 Two types of lattice stacking are observed in each of the two nanostructures (a) approximately 25 nm wide and approximately 2nm thick (b) approximately 10 mm wide and approximately 2nm thick. Different stacking are marked by arrows and are compared to the simulated images. These results show that all of the Tm silicide nanostructures have rectangular cross-sections, which is also consistent with the STM topography. Both the hexagonal Tm3$i5 and the orthorhombic TmSi are observed to grow epitaxially on Si(001). For the hexagonal ngsis, only the lattice parameter c was measured, but it did not vary from the 74 bulk phase. Nevertheless, for the orthorhombic TmSi phase, the lattice parameters vary fi'om the bulk to a significant extent. The relatively high number of nanostructures consisting of two phases, with the hexagonal phase overlying the orthorhombic phase, may represent a new mechanism for strain relief in this system. As was noted in section 2.2.2.1, the hexagonal and orthorhombic phases have very different compositions, which is different than for Gd silicide, and for most of the other RE silicide systems. Therefore, the coexistence of two silicide phases means the stabilization of an extreme compositional gradient within this nanostructure. Furthermore, the measured crystal structures imply that the orthorhombic phase richer in Tm lies at the interface with the Si substrate. If we consider the 2D lattice mismatch with the substrate, the hexagonal phase has 4%, and the orthorhombic phase has 7%, based on the bulk silicide lattice constants. However, the fact that the orthorhombic phase lies at the interface with the Si substrate, although it is richer in Tm, may suggest that it has a relatively small strain with Si(001). In fact the 2D lattice mismatch based on the measured lattice constants for the orthorhombic phase is 3% which is less than the hexagonal phase. Since there is poor contrast between the two phases in these structures, no clearly resolved mismatch can be observed. Nevertheless, this overgrowth pattern may suggest that the orthorhombic phase has a smaller 2D mismatch with Si substrate, while the hexagonal phase is more stable than the orthorhombic phase in term of volume free energy. 75 4.3 EELS Study of Gd Silicide Nanostructures To study the electronic nature of the Gd silicide nanostructures, the metallic Gd, thin film Gd Silicides, and Gd oxides in bulk form were studied systematically as references. Comparing the near-edge fine structures provides a way to investigate how the electronic nature of the nanostructures varies from the bulk phases. Both the N45 and M45 edges are detectable edges in the Gd EELS spectrum. Both of these edges represent transitions from initial (1 levels (respectively 4d and 3d) to final unoccupied 4f orbitals. Figure 4.17 shows the energy loss near-edge structure (ELNES) of the Gd N edge (transition of 4d to 40 fi'om the metallic Gd (a), the bulk GdSiz (b), the GdSi; nanostructure (c) and the Gd203 ((1) powder in comparison to Gd203 spectrum published in the standard EELS Atlas (e) [27]. The position and shape of the N peaks are similar in the four spectra, except for the presence of the Si L edge in bulk GdSiz (b) and GdSi2 nanostructure (0). Comparison of these spectra shows no difference between Gd oxide and metallic Gd in terms of ELNES of the N edge. Figure 4.18 shows the ELNES of the Gd M45 edge from metallic Gd (a), bulk GdSiz (b), GdSi2 nanostructure (c) and Gd203 powder ((1) in comparison to published data for Gd203 (e) [27]. Two peaks, M4 and M5, are well separated and correspond to unoccupied 4f states with different angular moments (M5 : 3d5/2—-> 4f, M4 : 3d3;2—» 40. Each experimental spectrum has been fitted using a power-law background, Hartree- Slater cross section and a set of Lorentzian profiles. The line with open squares demonstrates a good curve fitting result. The M peak positions and intensity ratios of the white lines are measured and listed in table 4.3. All four of the spectra were similar in 76 character. Within an energy resolution of 1.5 eV obtained from the full width of half maximum (FWHM) of the EELS zero loss peak, there are no measurable chemical shifts. Figure 4.19 shows the Gd electron configuration for the ground state neutral gaseous atom. It shows that 4f and 5d electrons have the highest energy states (65 has a lower energy state). In the metal, Gd has the nominal valence occupancy of 6s2 5d1 hence it tends to be trivalent in compounds such as Gd oxide. Nelson and coworkers state that the Gd in GdCOB has a 4f peak maximum located at 9.2 eV below the valence band edge whereas Gd 5d electrons contribute at the crystal valence band edge [57]. Photoemission measurements of Gd metal, Gd on silicon, and Gd3Si5 also show strongly localized 4f peaks about the same binding energy, whereas all states with about 5 eV of the Fermi level are either 5d or 63 derived (with some rehybridization with Si valence states[5 8]. Inverse photoemission measurements show that unoccupied 4f states lie roughly between 4 and 5 eV above the Fermi level [59]. Since Gd in nontrivalent states is not chemically stable [25], comparing these fine structures combined with the spectra of the other lanthanides in literature with various valences can give some indication of the Gd valence states. For instance, Thole and coworkers [60] have observed a chemical shifi of approximately 3eV between divalent Eu and trivalent Eu. Similarly, an approximately 4eV in chemical shifi has been observed in Tb3+ and Tb4+ [60, 61]. Therefore, the similarity of the M edge in the metallic Gd with the Gd oxide found in the present result confirms the assumption that the lanthanides are trivalent in the metallic solid state [62]. The similarity of the bulk silicides and the Gd oxide spectra is expected since Gd is also trivalent in the silicide form [63]. 77 The slight difference of intensity ratio between M5 and M4 in the GdSi; nanostructures with respect to the silicide is not readily explainable. Effects such as a surface shift in the unoccupied 4f states, which is seen in the metal and might be present in the silicide, might affect peak position but not branching ratio [59]. Any selection rule that might be present between 3d state and the multiplet structure of the empty 4f states would not easily translate into a branching ratio that varies with the chemical environment of the Gd atoms. GdN (a) Metallic Gd 3; (b) Bulk GdSi2 3° . a \ g) GdSi2 3.? /Si L anostructure E 0 E (d) Gd203 powder (e) Gdzos [27] a 1 L 1 FL I 1 i I I :41 r l r l n l r 100 110 120 130 140 150 160 170 180 190 200 Energy loss (eV) Figure 4.17 The energy loss near-edge structure (ELNES) of the Gd N edge from metallic Gd, bulk GdSiz, and Gd203 powder in comparison to the Gd203 spectrum from the standard EELS Atlas [27]. 78 (a) Metallic Gd (b) Bulk GdSi2 j "- (c) GdSi2 nanostructure Intensity (arb. units) (d) GdZOJ powder (e) Gdzo3 [EELS Atlas] l l L I 1 l n l I 1 I a l 1150 1160 1170 1180 1190 1200 1210 1220 1230 1240 Energy loss (eV) Figure 4.18 The energy loss near-edge structure (ELNES) of the Gd M4; edge fiom metallic Gd, bulk GdSiz, and Gd203 powder in comparison to the Gd203 spectrum from the standard EELS Atlas [27]. The line with open squares demonstrate good curve fitting using a power—law background, Hartree-Slater cross section and a set of Lorentzian profiles. 79 4f D ID 1D ID ID ID 1E Dll H—lflfl 5d IEE 5 ”mam“ " [EII 4P 3d tifllEEEE 4s 3p IEI 3S 2pIIlE 25 E Is. 55 '65 Figure 4.19 Gd electron configurations for the ground state neutral gaseous atom. Table 4.3 Summary of N and M peak positions and intensity ratio measurements of the white lines fiom the metallic Gd, the bulk GdSiz, the GdSi; nanostructure (NS), and Gd203 powder in comparison to Gd203 fiom EELS Atlas [27]. . . Branchin Peak posrtron (eV) ratio g N M5 M4 Ms-M4 Ms/M4 Metallic Gd 149.3 1186.4 1214.6 28.2 2.02i0.06 Bulk GdSi2 148.9 1186.1 1214.6 28.5 1.97i0.02 GdSi; NS 148.7 1186.0 1215.5 29.8 1.74'_1'0.02 Gd203 powder 149 1186.2 1215.5 29 2.03i0.02 Gd203 149 1188 [27] 1217 [27] 29 1.73 [26] 4.4 Conclusions HRTEM was used to characterize epitaxial silicide nanostructures of two different RE metals: Gd and Tm. In the case of Gd, the silicide nanostructures are either the hexagonal or the orthorhombic phase, consistent with prior studies. However, some bi- phasic silicide structures are seen, with layers of the orthorhombic phase overlying the hexagonal. In the case of Tm, the nanostructures are either hexagonal or orthorhombic. Once again, some nanostructures are seen with one silicide phase overlying the other. For Tm, the orthorhombic phase lies at the interface between the hexagonal and the orthorhombic. For both systems, bi-phasic silicide structures may reflect a mechanism for strain accommodation at the interface with the substrate. In the case of Gd, the phase with lower strain lies at the substrate. For the case of Tm, the relative mismatches of the two phases predicted from bulk silicide lattice parameters disagree with that derived from measured lattice constants, and it is a relaxed orthorhombic phase at the interface that appears to have the lowest mismatch with the substrate. This could also reflect the fact that this interfacial phase is only two or three atomic layers of Tm silicide thick. EELS studies were carried out to compare the electronic structures of metallic Gd, thin film Gd silicide, and Gd oxide in bulk phases and Gd silicide nanostructures. The results fi'om the three bulk phases are similar, while the intensity ratio of M5:M4 in the GdSi2 nanostructures varies from the bulk, which may suggest that a slightly different spin state exists in the silicide nanostructures. 81 Chapter 5 Results from Multilayer Structure Studies This chapter presents the HRTEM and EELS results from the various metallic multilayers examined in this study. The first section will cover the Co/Al and Py/Al multilayers, followed by the results from the Cu/Al multilayers and Cu/Al in the Py- based spin-valve structures. 5.1 Co(Py)/Al Multilayers The Co(Py)/Al multilayers were composed of ten Co(or Py)/Al bilayers with a 20 nm Co (or Py) layer and a 10 nm Al layer sputtered on Si(lOO) substrates at room temperature. To compare with the as-sputtered samples, some multilayers were annealed for less than 5 min after sputtering at 180°C in the same chamber. The results fiom Co/Al and Py/Al multilayers are presented separately. 5.1.1 As-sputtered and Annealed Co/Al Multilayers A phase contrast cross-sectional TEM image of the as-sputtered Co/Al multilayers on Si is shown in figure 5.1. The well-layered Co and A1 are both polycrystalline. The thicknesses of the individual layers are consistent with the nominal 20 nm Co and 10 nm A1, with in-plane grains sizes in the range of ~20 nm. Significant layer roughness, in the form of non-planar interface perturbations, is observed, resulting in the interfacial planes lying as much as 15° from the growth plane. A selected area diffraction (SAD) pattern fi'om both the multilayers and Si substrates (see inset) shows strong diffraction from the close-packed planes of the Co and Al in the growth direction, which indicate the layers are highly textured. 82 7 _ Si(OOZ) ‘0 O ‘Q‘Amm Cotilll) Figure 5.1 Phase contrast TEM image of as-sputtered Co/Al multilayers on a Si substrate. Inset is a selected area diffraction pattern from both the multilayers and Si substrates showing strong diffiaction fiom the close-packed growth planes of both the Co and Al, parallel with Si(001). In the upper right, an inclined grain is illustrated schematically to show that the waviness perpendicular to the electron beam is up to 2 nm at the interfaces. It is well known that Co can exist as either FCC or HCP [19]. In many cases [64] these two phases effectively co-exist as heavily stacking faulted permutations of either perfect structure, which results from the low stacking fault energy that in turn results 83 fiom the very small difference in formation energy of these allotropes [65]. Stacking faults can influence the magnetic behavior of Co, with higher fault densities decreasing the MR [39]. Thus, it is important to characterize the crystal structures and faulting of the Co in the multilayers in this study. Figure 5.2a shows a Co/Al/Co HRTEM image in which the Co and Al layers are clearly evident. Two Al grains are clearly resolved in the light contrast layer, revealing an ~7° angle between the two Al(111) growth planes. Within the Co layers, the atomic structure is resolved, but somewhat obscured by moiré fringes, which are caused by overlapping grains‘in the beam direction. The FFT of the image (see inset) shows two strong Al diffraction spots fiom the {111} close-pack planes of the two A1 grains, but a variety of FCC and HCP Co diffi‘action spots. Since the image is not taken fi'om a well-aligned Co crystal zone (as indicated by the FFT), the Co close- packed plane stacking is not clearly defined. Figure 5.2b shows another area from this multilayer with both the Al and Co clearly resolved. Most of the Co in this image displays a <110>FCC zone axis with ABCABC stacking of the close-packed {111} planes (see upper inset). However, some HCP Co, with ABAB stacking (see lower inset), is also observed. Analysis from different areas in the multilayers shows that FCC stacking is more dominant than HCP stacking. An interesting feature of these multilayers is that although both the Achc and Com: grow on their close-packed {111} planes, Al and Co do not form coherent {111} interfaces. Instead, a significant angular deviation of ~7° between the {111}A1 and {111}co plane normals is found, as illustrated in the FFT inset in figure 5.2b. Although this was most evident in grains imaged in the <110> orientation, such deviations were consistently observed along the interfaces. 84 It is noted that the interfaces of Co/Al layers are obscure in figure 5.2, which makes it difficult to identify any inter-metallic phases that may have formed by diffusion at the interfaces. A through-focus series of images in figure 5.3 can help illustrate the interfacial structures along the electron beam. The four images were taken from the same area, but at defocuses of 0, 32, 64, and 96 nm. For a defocus of 0, the Al lattice fiinges extend through the interfacial region and to the Co at the upper interface. As the defocus is changed, this distinct stacking of Al planes becomes obscured at the upper interface, but extends closer to the lower interface, until it finally reaches the Co layer at a defocus of 96 nm. HRTEM image simulations have shown that at different defocuses lattice fiinges from a Ni3Al crystal will terminate at different positions relative to an inclined grain boundary [66, 67]. Thus, this through-focus series suggests that what appears to be an interfacial region between the Co and Al is actually the result of the interfaces being inclined to the electron beam direction. Given the observed interface and layer roughness perpendicular to the electron beam of approximately 2 am over a width of 20 nm illustrated in figure 5.1, one would expect the interfaces to also exhibit some waviness in the electron beam direction. With a TEM foil thickness of 10 nm and an observed interfacial region thickness of 1 nm, the interfacial waviness in the electron beam direction is consistent with the interfacial waviness perpendicular to the electron beam direction. Figure 5.4 shows a schematic model of Al{111} growing on Co{111}. The mismatch between the two lattices, calculated based on the lattice parameters of bulk FCC Al and Co listed in Table 5.1, is 13.7% for A1{1 1 1} growing on Co{111}. Clearly, it is unlikely that these layers will directly hetero epitax on each other [1 l]. 85 ,Co ( ,‘ ll Figure 5.2 HRTEM images showing Co/Al atomic lattices with different contrast. The Co/Al interfaces are poorly defined. A grain boundary in Al is observed in (a). Both FCC and HCP stacking in the Co are shown in the enlarged images in (b). Insets in the upper right show the corresponding FFTS from the images. 86 Defocus 0 CO Defocus 32 Defocus 64 Figure 5.3 A through-focus series of images from the same area of the multilayer indicates that the Al and Co interfaces are inclined along the beam direction. Instead, it appears the mismatch is accommodated by significant tilting between those lattices, as shown in figure 5.4, which results in the inclined interfaces and apparent roughness. The resulting interfaces would be expected to be complex, with atomic ledges and dislocations with Burger vector both in and normal to the interface. 87 Figure 5.4 Schematic growth model of FCC A1 {111} on FCC Co {111} with a lattice mismatch of 13.7%. Table 5.1 Lattice parameters and close-packed interplaner spacing variants for the multilayers. Materials Lattice parameters Interplaner spacings Interplaner spacings for (nm) measured for (111) (111) in theory Al ao= 0.405 0.235i0.002 mm 0.234 nm ao=0.251, Co co=0.407 (HCP) N/A N/A ao=0.356 (FCC) 0.205i0.002 nm 0.206 nm Py (NiFe) ao= 0.355 0.204i0.002 mm 0.205 nm To study the effect of annealing on the Co/Al multilayers, both TEM and spatially *dSiooz= 0.272nm is used as reference resolved EELS were carried out on both the as-sputtered and annealed samples to study the structures and chemical nature of the Co/Al interfacial region. Figure 5.5 shows a phase contrast TEM image of an annealed Co/Al multilayer. The morphology and 88 corresponding SAD pattern demonstrate that the layers maintain well-defined Co and Al layers and no obvious intermediate phases. 0 Si (002') %‘AHHH XI‘OUH) Figure 5.5 Phase contrast TEM image of an annealed Co/Al multilayer. Inset is a selected area diffraction pattern from both the multilayers and Si substrates showing strong diffraction from the close-packed growth planes of both the Co and A1. Figure 5.6a shows a high angle annular dark field (HAADF) STEM image of the as-sputtered Co/Al multilayers. The image contrast is proportional to Z2, so the Co (Z=27) layers are much brighter than the Al (Z=13) layers. As a result, the analysis of 89 the image contrast can give some indication of the compositional information at the interfacial region by comparing the as-sputtered and annealed samples. A comparison of line profiles of image grey level across a Co/Al/Co layers fi'om both the as-sputtered and annealed multilayers is shown in figure 5.6b. Although some differences in image grey levels in the Co and A1 layers are shown, there are no significant differences in terms of contrast change in interfacial regions. A series of EELS spectra were collected by line scanning the electron probe across layers. Figure 5.6c shows a perspective view of an EELS spectrum-image with 61 regularly spaced low loss spectra obtained along the 49 nm line indicated in figure 5.63. The square frame labeled “Spatial Drift” indicates the area used as the reference to correct the specimen drift during the spectra collection. 90 Spectrum Image (b) 250 _ annoaied 200 ~ 150 > 7a 5 3 100 . , as-sputtered Spatial (3 Drift Co AI 50 » c° t o I l ‘ l I l 0 5 10 15 20 25 30 35 40 A1 plasmon Distance (nm) Energy loss (e ' Figure 5.6 (a) A dark field STEM image of the as-sputtered Co/Al multilayers showing where the EELS spectrum image was acquired along the white dashed line. The spatial drift square is the cross-correlation scan area used to correct the drift. (b) A comparison of line profiles of image grey level across a Co/Al/Co layers from both as-sputtered and annealed multilayers. (c) A perspective view of an EELS spectrum image showing the plasmon intensity changing across the Co/Al multilayers. It is known that Al has a very intense plasmon peak compared to many metals, including Co, Fe, and Ni [27]. Thus, the low loss spectrum of plasmon intensities can give an indication of the elemental distribution across the spectrum image. Figure 5.7 shows a comparison of the plasmon intensity profiles for the as-sputtered and annealed 91 Co/Al multilayers. Both of the plasmon peak profiles across the Al-rich layers are 5-6 am in full width at half maximum (FWHM), which indicate similar Al distribution in the as-sputtered and annealed samples. Within the Co-rich layers, relatively constant regions, approximately 15 nm wide for the as-sputtered multilayer and 17 nm wide for the annealed multilayer, suggests there are no significant differences in the Co distribution in the two conditions. Normalized Intensity 1.0- 1.0- 0.8 E’ 0.8 - g . 0.6 g 0.6- 0.4 '7. 0.4- F. O 0.2 2 0.2 0.0...1 ................... 0.01:. .I.L.I.LII.I.I. . 04812182024283236404448 048121620242832364044 Figure 5.7 A comparison of the plasmon intensity profile for as-sputtered and annealed Co/Al multilayers. The Al plasmon peak profiles have full width at half maximum of 5-6 nm, similarly for both as-received and annealed samples. The only difference is that the low variation of Co intensity length is 15 nm for the as-received multilayers and 17 nm for annealed one. An alternative way to characterize interfacial compositions using EELS is to investigate the Co L23 edge intensity profile across Co/Al interface, which can provide sufficient insight in the chemical nature at the Co/Al interfacial region. Figure 5.8 shows a perspective view of the EELS spectrum image recorded around the Co L23 edge (L3z779eV, L22794) across the Co/Al/Co layers of the as—sputtered sample. Upon removing a power law background for each spectrum, the intensities of L3 peaks were 92 analyzed and plotted in figure 5.9 (shown as open circles). The intensity profile is compared to a similarly processed intensity profile for the annealed layers (solid squares in figure 5.9). To facilitate an understanding of the two experimental profiles, a theoretical Co intensity profile across perfect Co/Al/Co interfaces (see in figure 5.9) has been simulated using an electron beam with a 0.5 nm (FWHM) Gaussian profile. The experimental data reveals significantly broader interfacial mixing than the theoretical profile. While some of this broadening will be associated with the inclination of the interfaces described earlier, and possibly from beam spreading in the TEM thin foil (both ignored in the simulation),the deviation from the theoretical profile may suggest some mixing across the interfaces. Nevertheless, the intermixing is not sufficient to initiate additional phase formation. 93 .th \tff' tttw 42111:“ \p‘llq‘l ['9 N it] “ii?!“ 1.1, ,, l ltidy“ ""1'1‘”'stii'a'it'ktiii ' a» ti ‘52:le {15% 1M") I' , . “‘1‘ ll” W \.' 13.11:} w! ( “\ ”all!“ fill}; \ ll ‘ “51‘1” ‘5'"; \tttt ‘~'~' .-.*~I t‘l‘tl . \l.‘ .u,’ (,‘v immmmw all Probe position (nm) Figure 5.8 A perspective view of the Co white line spectrum image scanned across Co/Al/Co layers over a length of 20 nm of the as-sputtered multilayers. Normalized EELS Intensity Comparison of Co Intensity Profiles I r I ' I ' l ' l ' I ... Annealed data — Simulation curve C39 0 00 O As—sputtered data l I l 0 2 4 6 8 10 12 l4 16 Distance lnml 18 20 Figure 5 .9 Co white line intensity profiles for both the as-sputtered and annealed multilayers. A theoretical profile simulated by using a 0.5nm FWHM Gaussian electron beam profile across perfect Co/Al interfaces is used to compare with the experimental data. 94 5.2 As-sputtered and Annealed Py/Al Multilayers This section presents HRTEM and EELS results fiom the Py/Al multilayers sputtered on Si(lOO) substrates. The multilayer was composed of ten bilayers with a 20 nm Py layer and a 10 mm Al layer Phase contrast TEM images of both as-sputtered and annealed Py/Al multilayers grown on Si are shown in figure 5.10. Py (permalloy = Nil- xFex with x~0.2) has an L12 fcc crystal structure. The images reveal the layers are polycrystalline, while the strong diffraction peaks for the close-packed Py and A1 {111} planes in the growth direction in the SAD patterns (see the insets in figure 5 .10) demonstrate that Py and A1 are highly textured. The thicknesses of the individual layers are consistent with the nominal 20 nm Py and 10 nm Al. The Py had grains with sizes in the range of ~20 nm, and similar to the Co/Al multilayers, some of the interface normals are tilted. The observed structures and interface roughness do not reveal any obvious differences between the as as-sputtered and annealed samples. Figure 5.1] shows an HRTEM image of an annealed Py/Al multilayer with most of the Py and some of the Al clearly resolved. ABCABC stacking of the close-packed {111}pcc planes are displayed in both Py and Al. The FFT (see lower inset) shows that the grains are imaged along <110>pcc zones of Py and A]. Again, a significant angular deviation of 7° between the {1 11}A1 and {11 1}py plane normals is found, which will lead to complex interfaces. Similar to the observation in the Co/Al multilayers, the lattice fringes are obscured near the interfaces of the Py/Al layers in the image, which can be interpreted as resulting of inclined interfaces. 95 Py 002 : - Atooz’” , g . “ P), 1 11 o “» v. Ahl-H s: 002 Figure 5.10 Phase contrast TEM image of as-sputtered and annealed Py/Al multilayers. Insets are selected area diffraction patterns from the multilayers. Both morphologies and diffraction patterns show similar structures in two conditions. 96 Figure 5.11 HRTEM image of an annealed Py/Al multilayer along a Py <110> zone with corresponding FFT in the inset showing the growth relationship. It is known that both Fe and Ni can form a range of intermediate phases with Al, including several ternary intermediate phases [1?]. To study if there is any intermixing at the Py/Al interfacial region in the as-sputtered and annealed multilayers, spatially resolved EELS was carried out to study the composition across the interfaces. EELS spectrum images that included the low energy plasmon peaks and Ni and Fe L edges (Ni L328556V, Fe L3:708eV) have been collected across the interfaces for both the as- sputtered and annealed samples. Figure 5.12 shows a comparison of the normalized 97 Normalized Intensity plasmon intensity profiles of across the Py/Al/Py/Al/Py layers, which reveal layer thicknesses consistent with the nominal thicknesses. Figures 5.13 a and b show the comparisons of intensity profiles of Ni and Fe at the interfaces of the as-sputtered and annealed Py/Al/Py layers with the simulated theoretical profiles. Figure 5.13c shows the comparison of line profiles of the image grey level from across Py/Al/Py layers in the Z- contrast images of the two conditions. No significant differences were observed between the annealed and as-sputtered Py/AJ/Py profiles, which suggests the annealing treatments do not significantly alter the multilayer structure. Nevertheless, the intensity profiles do suggest that some limited compositional mixing may occur during the initial growth. As—sputtered Plasmon Loss Intensity Across Interfaces Annealed Plasmon Loss Intensity Across Interfaces l-o _ I I I I I l I I I I - I I I I I I ' I I I I I .- I‘ 1.0 _ II .- I I I I I - I I - 0-3 Py Al Py LA. I Pr =:‘ o s Py .Al. Py IAI- 1’! - ----- I---u-----------: -- -- 5 '"'i"'"""""""lL'fl'" 0 6 > I a . _ I 7nm l 23m l 7mm .— 0.6 _ . 6.5 nir- 22 5 um I 6.5m!- “ y I E I . o 4 '- T- o 4 # uh. I _ 2g 17 ‘ o 2 ' 16nm I ‘ 03- ' l ——-|nm ' 0.0 0.0 I | I l l I I l l I I l I I I I I I I I l I 0 4 8 l2 l6 20 24 28 32 36 40 44 48 52 0 4 8 12 16 20 74 28 32 36 40 44 48 52 Distance (nm) Distance (nm) Figure 5.12 A comparison of plasmon intensity profile from the as-sputtered and annealed Py/Al multilayers. The A1 plasmon peak profiles have full width at half maximum of 6.5-7 nm, similarly for both as-sputtered and annealed samples. 98 (3) Comparison of Ni Interface Profiles I 2 I I I T I I I E 1.0 — a E .5 0.8 m :3 H 0.6 E = 0.4 E g 0.2 0.0 I l l J n 1 a I; 0 4 8 I2 16 20 24 28 Distance (nm) (b) Comparison of Fe Interface Profiles I r I r Normalized EELS Intensity Distance (nm) (e) 250 - as- \ sputtered 200 — T; 150 — 2 g annealed (J 100 r- Py AI Py 50 e . 0 a l L J A a 4.. 1 L4L L I 0 4 8 l2 16 20 24 28 Distance (nm) Figure 5.13 Ni (3) and Fe (b) white line intensity profiles for as-sputtered and annealed multilayers are compared separately with the theoretical profile simulated by using a 0.5nm FWHM Gaussian electron beam profile across perfect Py/Al interfaces. (c) A comparison of line profiles of the image grey level from Py/Al/Py layers in the Z-contrast images of the two conditions. 99 5.2 Cu/Al Multilayer and Cu/Al Spin-Valve Structures To study the effects of Al interlayers on Py/Cu/Py based spin-valves, structures of alternating Cu/Al multilayers have been investigated by HRTEM. Since increasing the thickness of the Al layer in the spin-valves can cause a decrease in the resistance [48], different thickness ratios in Cu/Al multilayers were synthesized to see if the structures vary with different conditions. Multilayers made up of Cu(8nm)/Al(lOnm) with an overall composition of 52.9 at.% Cu and Cu(5nm)/Al(3nm) with an overall composition of 70.1% at. Cu were selected for HRTEM studies. These multilayers were examined in a number of post sputtering states, specifically as-sputtered and annealed, to investigate thermal instability of these multilayers. TEM samples in both states were prepared using specific approaches to minimize sample heating (described in 3.4). Results from HRTEM imaging with associated F FT analysis show that different intermediate phases form in the Cu/Al multilayers with different thickness ratios. In order to determine if the intermediate phase formation observed in multilayers also occurs in a single Al layer bounded by Cu in Py based spin-valves, selected spin- valves were characterized. The spin-valves have the following configuration, where all the values are given in nanometers: Nb(l50)Cu(5)FeMn(8)Py(24)Cu(10)Al(x)Cu(10)Py(24)Cu(5)Nb(50). In this study, spin- valves with Al(x=10) for as-sputtered and Al(x=30) for annealed were selected. The spin-valves were studied using HRTEM and EDX line scans to characterize the crystal structures and chemical nature. 100 5.2.1 Cu/Al Multilayers 5.2.1.1 Cu(8nm)/Al(10nm) Multilayer Structures A phase contrast cross-sectional TEM image of the as-sputtered Cu(8nm)/Al(10nm) multilayer is shown in figure 5.14. Well-defined alternating layers of different contrast are observed. The layers are polycrystalline and some moiré fiinges are observed. The Cu-rich layers exhibit dark contrast and are ~12 nm thick, while the Al- rich layers exhibit bright contrast and are ~6 nm thick. By comparing these measured layer thicknesses to the nominal, it appears that more Al has diffused into the Cu rather than the opposite. In addition, this change in layer thickness ratio suggests that intermediate phase formation may have occurred. Cu rich layer Al rich layer Growth direction 20 nm Figure 5.14 Phase contrast TEM image of as-sputtered Cu(8nm)/Al(10nm) multilayers. Observed layers are in different contrast, exhibiting ~12nm thick Cu- rich layer and ~6nm thick Al-rich layer, which are different from the nominal thickness. HRTEM imaging with associated FFT analysis were can'ied out to characterize the structures of the intermixed Cu and Al. Figure 5.15 shows an HRTEM image of the 101 same Cu/Al multilayer. Five layers are shown in the image: Al-rich layers labeled 1, 3, and 5; Cu-rich layers labeled 2 and 4. This image show different fringe contrast and symmetries in the different Al-rich layers. The Al-rich layers are identified to be tetragonal AlzCu (lattice parameters a= =0.604nm, c=0.486nm). The FFTS (see insets) from the Al-rich layers 1 and 5 show the same A12Cu <001> zone pattern, while the FFT from Al-rich layer 3 shows an A12Cu <11 l> zone pattern. All the lattice spacings measeasured are summarized in table 5.2. The FFTS from the Cu-rich layers (2 and 4) show three-fold symmetry, which suggests the phase has a cubic structure. However, the lattice spacings are different (by ~60%) than that expected for fee Cu. Instead, the F FT pattern is consistent with the <111> zone axis for the bee [3 solid solution of AICU3. The [3 phase is known to exist at high temperature in the Cu-Al system, and the [3 phase can be retained as a metastable phase due to the sluggishness of the eutectoid reaction B—+ Cu + 71 [17]. Both the A12Cu and AlCu3 are orientated with their close-packed planes of (110) and (110), respectively, normal to the growth direction. Based on the FFT analysis, two types of growth relationships are found: A12Cu(110) // AlCu3(110), A12Cu[001] // AlCu3[lll], and A12Cu(110)//A1Cu3(110), A12Cu[lll] //A1Cu3[111]. 102 0 AIZCir <111> " B AlCu, Figure 5.15 HRTEM image shows three types of stacking fringes with different contrast existing in the Cu/Al multilayers. Five different areas in the image are analyzed using FFTS, resulting in three crystal zone patterns shown in the insets. Areas 1 and 5 show the same 9 A12Cu <001> zone patterns, areas 2 and 4 show the same B AlCu3 <111> zone patterns, and area 3 shows a 9 A12Cu <001> zone pattern. 103 Table 5.2 Summary of the lattice spacings measured from all the phases in the Cu/Al multilayers. Theoretical Measured Layer Plane . . spacrng (nm) spacmg (nm) (110) 0.427 0.42 A12Cu (011) 0.378 0.38 (020) 0.302 0.30 (112) 0.211 0.21 AlCu3(bcc) (1 10) 0.205 0.20 1 1 1 0.209 0.21 Cu(fcc) ( ) (002) 0.181 0.18 Figure 5.16 shows a schematic representation of the two phases and their growth orientation relationships. Both bcc AICU3 (figure 5.16a) and tetragonal AlZCu (figure 5.16b) will grow on their close-packed {110} planes, but these planes can be aligned in two different directions. The dashed rectangle in figure 5.16c shows a structural unit in the AlCu3(110) plane that can line up in two different ways on the A12Cu, as shown in figure 5.16d. When AICU3[111] // A12Cu[001] (the fine dotted lines on figure 5.16d), there is a very small mismatch (0.5%) in one direction and a large mismatch (-16.6%) in the perpendicular direction. When AlCu3[l l 1] // A12Cu[l 1 1] (the coarse dotted lines on figure 5.16d), there is a relatively small mismatch (2.4%) in one direction and a even larger mismatch (-1 7.6%) in the perpendicular direction. Consequently, there are two rotation alternatives for AlCu3 growing on AlzCu. This can result in a variety of grain orientations throughout a columnar grain. To study the thermal stability of these Cu/Al multilayers, some of the Cu(8nm)/Al(10nm) multilayers were annealed at 180°C for 5 min. Figure 5.17 shows a 104 phase contrast TEM image of the annealed Cu(8nm)/Al(10nm) multilayers. The Cu-rich layers exhibit dark contrast and are ~13 nm thick, while the Al-rich layers exhibit bright contrast and are ~7 nm thick. The layer thicknesses are different with the nominal, but similar with those of the as-sputtered multilayers. Again, the Cu-rich layers are thicker than the nominal and the Al-rich layers are thinner. This suggests more diffusion of the Al into the Cu than the Cu in the Al, and possible intermediate phase formation in the multilayers. An I-[RTEM image from these annealed Cu/Al multilayers is shown in figure 5.18. The FFT analysis again shows that the Cu—rich phase is AlCu3 and the Al- rich phase is A12Cu. All of the measured lattice spacings are summarized in table 5.2. In conclusion, similar structures with the same intermediate phases are observed in the as- sputtered and annealed Cu(8nm)/Al(10nm) multilayers. The observed difference is that the interfaces are wavier in the annealed multilayer compared to those in the as—sputtered multilayer. 105 (a) AlCu3 crystal structure (b) AlzCu crystal structure 0.243nm (-l 7.6%) *_ __ vow 9V 0.427nm (2.4%) fl AlCu3 [111] (c) AlCu3 (110) (d) A12Cu (110) Figure 5.16 Schematic representation of the bee AICU3 and tetragonal A12Cu crystal structures and the orientation relationships of their close-packed planes. (a) bcc AICU3 crystal structure (b) tetragonal A12Cu crystal structure (c) a structural rectangle unit in AlCu3(110) (d) two orientation relationships with AlzCu(110): AlCu3[11 1] // A12Cu[001] (the fine dotted lines) and AlCu3[l l 1] // A12Cu[111] (the coarse dotted lines). 106 « «~er %“W «We. Aficmhld30£am® . Wfimnm Growth direction Figure 5.17 Phase contrast TEM image of annealed Cu(8nm)/Al(10nm) multilayers. The Cu-rich layers are ~13 nm thick and the Al-rich layers are ~7 nm thick. AIZ'Cu(-l10) Cu—rich .AlClrj(-l,l(|) Figure 5.18 HRTEM image with corresponding FFTS from the annealed Cu(8nm)/Al(10nm) multilayers. The Al-rich phase and the Cu-rich phase are identified to be A12Cu and AlCu3 respectively. 107 5.2.1.2 Cu(5nm)/Al(3nm) Multilayer Structures Figure 5.19 shows a phase contrast TEM image of the as-sputtered Cu(5nm)/Al(3nm) multilayers. The Cu-rich layers are 34 nm thick, while the Al-rich layers are 4-5 nm thick. The layer thicknesses are different with the nominal, which again suggests that intermixing, and possibly intermediate phase formation, have occurred. An HRTEM image from these Cu/Al multilayers is shown in figure 5.20. FFTS (insets) from 4 areas were used to identify the phases. The FFTS from the Al-rich layers 1 and 3 show the same AlzCu <111> zone pattern, while F FTs from Al-rich layer 2 and 4 show the same FCC Cu <110> zone pattern. All the lattice spacings measured are surmnarized in table 5.2. The growth orientation relationship is identified to be AlzCu(110) // Cu(l 1 1), AlzCu[1 11] // Cu[1 10]. Compared with the Cu(8nm)/Al(10nm) multilayers, Cu did not form a bee solid solution with Al in the Cu-rich layer in the Cu(5nm)/Al(3nm) multilayers. This may result from the difference in the overall stoichiometry of the Cu and Al, but the reasons are not entirely clear. Figure 5.21 shows a schematic representation of the close packed planes of the two phases, fcc Cu (111) and tetragonal A12Cu (110), and the way they epitax with each other. There is an equilateral triangular unit existing in the Cu (111) planes and a similarly orientated isosceles triangular unit existing in the AlgCu (110) planes. Matching the two units results in a -4% mismatch in two directions and a -5% mismatch in the third direction. When Cu is epitaxed on to the AlzCu, there is effectively only one 108 Figure 5.19 Phase contrast TEM image of the as—sputtered Cu(5nm)/Al(3nm) multilayers. The Cu-rich layers are 3-4nm thick and the Al-rich layers are 4-5 mm thick. ..., . '.';¢ N -‘-’II!;:;;.¢I . t . .,.-otvaaov ....~aetv-:¢- Figure 5.20 HRTEM image shows two types of lattice fringes existing in the Cu(5nm)/Al(3nm) multilayers. Four different areas in the image are analyzed using FFTS, resulting two crystal zone patterns shown in the insets. Areas 1 and 3 show the same 0 A12Cu <111> zone patterns, while areas 2 and 4 show the same Cu <110> zone patterns. 109 orientation that the Cu can assume. However, the AlzCu can align in three different ways on the Cu as illustrated in figure 5.21. This may result in significant layer to layer orientation variation in a columnar multilayer grain. ,0. t7". —» I c [110) “ .----QW, FCCCu(111) A12Cu (110) Figure 5.21 Schematic representation of FCC Cu (111) and tetragonal Alzcu (110) growth orientation and lattice mismatches. 5.2.2 Cu/Al in Spin-Valve Structures 5.2.2.1 As-Sputtered Spin-Valve with 10nm Al Two different Cu/Al spin-valve structures were selected for study. They have the following configurations, where all the values are given in nanometers: Nb(150)C\1(5)FeMn(8)Py(24)Cu(l0)Al(x)Cu(10)Py(24)Cu(5)Nb(50), In this study, a sample with Al(x=10) was examined in the as-sputtered state and a second sample with Al(x=30) was examined in the annealed state. The spin-valves were characterized using HRTEM and EDX line scans to determine their crystal structures and chemical nature. 110 While this study focused on the Cu/Al layer structures, other layers in the spin valves were imaged and used as references (i.e. Py atomic structures were used as reference to measure the lattice spacings). Figure 5.22 shows a TEM bright field image of all the layers in the as-sputtered spin-valve with the nominal configuration with Al (x=10nm). The thickness of some of the individual layers is not clearly resolved in this bright field image. In fact, the Cu/Al/Cu portion of the spin valve appears as one layer. Nevertheless, the image does show that the spin valve is layered with polycrystalline grains. Twins and stacking faults are evident in some of the grains. Figure 5.23 shows a phase contrast TEM image of the PyCuAl region with only the Al layer in clear contrast. While it is difficult to measure the thickness of the Cu and Py layers due to the similar Z, some region shows the ~10 nm thick Cu layer, marked by the arrows in the image. The observed thickness of the resolved layers is consistent with the nominal. To study the crystal structures of the Cu/Al layers, Py and Nb, which have known structures, were first studied to establish references. Figure 5.24 shows an HRTEM image (a) of Py/FeMn/Cu/Nb layers in the as-sputtered spin-valve and the corresponding FFT (b) from the image. Moiré fringes, caused by overlapping grains at the interfacial region, make it difficult to resolve the FeMn and Cu structures. However, the Py and Nb layers show clear atomic lattices and the FFT (b) shows strong diffraction from the Py(l 11) and Nb(110), which were used to as reference to measure the lattice spacings (listed in table 5.3). 111 §_ \- I. Cu/Alglrguexs. ’ Py Cu, FeMn\ 50mm -—"' Si substrate Figure 5.22 TEM bright field image of an as-sputtered spin-valve with nominal thickness in nanometers Nb(l 50)Cu(5)FeMn(8)Py(24)Cu(l0)Al(10)Cu(l0)Py(24)Cu(5)Nb(50). 112 Figure 5.23 Phase contrast TEM image of Py/Cu/Al/Cu/Py layers in the as-sputtered spin-valve with x=10nm. 113 FeMn/Cu ‘4 ..- ."t" (h) P}(lll) X|l(llll) Figure 5.24 (a) HRTEM image of Py/FeMn/Cu/Nb layers in the as-sputtered spin-valve with x=10nm and (b) the corresponding FFT from the image. 114 Table 5.3 Summary of the lattice spacings measured from all the phases in the as-sputtered spin-valve. Layer Plane Theoretical Measured structure spacmg (nm) spacmg (nm) Py (l l 1) 0.205 As reference Nb (110) 0.233 0.23 Cu(fcc) (l 1 1) 0.205 0.2 AlzCu (220) 0.214 O. 22 A1 (111) 0.234 N/A Figure 5.25 shows an HRTEM image with the Cu and Al region exhibiting different contrast. The corresponding FFT shows two diffraction peaks along the grth direction, very close to each other, and diffraction at half the distance in the reciprocal space, which suggest a structure with a superlattice formed. This observation is consistent with the double periodicity observed in the Al-rich region of the HRTEM image. These observations suggest an intermediate phase has formed. Indeed, measurements on the FFT shows that the closely spaced diffraction peaks are from CU(111)f¢c and AlzCu(220), (see table 5 .1). This analysis is consistent with that carried out in other Cu/Al regions of the spin valve. No fee Al was observed in the Cu/Al region throughout the multilayer. Complementary XEDS line scan were carried out to investigate the chemical nature across the multilayer structure. Figure 5.26a shows a STEM bright field image of the as-sputtered spin-valve. XEDS intensity profiles of Cu, Ni, Fe, and Al were collected from the line across the Py(24)/Cu(10)/Al(lO)/Cu(10)/Py(24)/FeMn(8)Cu(5) 115 1' "111.; a ti- , A ' ’2“. 3 Al melt in “lift“. ‘~ (b) Al,Cu(220) it A Cu(111) Figure 5.25 (a) HRTEM image of Cu/Al/Cu layers in the as-sputtered spin-valve with x=10nm and (b) corresponding FFT from the image. 116 o 50 100 nm COppCfKfllJiukrrlbéiL ' ~ 1‘ AlumnmmKat Figure 5.26 (a) STEM bright field image of the multilayers in the as-sputtered spin- valve. (b) XEDS intensity profiles of Cu, Ni, Fe, and A] were collected from the line across the Py/Cu/Al/Cu/Py/FeMn/Cu layers in the image (a). 117 layers. The FWHM for adjacent Py, Cu and Al layers are 25, 12, and 14 nm respectively. I The line profiles indicate that there may be some intermixing taking place between the Cu and Al, while the other regions remain well segregated in their metallic structures (i.e. the measured FWHM of 25nm for Ni and 6nm for Cu compare very well with their nominal thickness). Both HRTEM and XEDS analyses suggest that there is some intermixing in the Cu/Al region in the as-sputtered sample with Al 10nm. 5.2.2.2 Annealed Spin-Valve with 30nm Al Figure 5.27 shows a TEM bright field image of all the metallic layers in the annealed spin-valve with Al(x=30), i.e. with a nominal configuration of Nb(150)Cu(5)FeMn(8)Py(24)Cu(l0)Al(30)Cu(l0)Py(24)Cu(5)Nb(50). All the layer structures are polycrystalline. The thickness of some of the individual layers is not clearly resolved in this bright field image. However, the Al layer shows clear bright contrast compared to the adjacent Cu layers. Figure 5.28 shows a phase contrast TEM image of the PyCuAl region that shows the layer more clearly resolved. The Py layers display a thickness of approximately 24nm consistent with the nominal. However, the Cu-rich layers, approximately 14nm thick, are significantly thicker than the nominal, while the Al-rich layer is much thinner, approximately 18nm thick, than the nominal thickness. This again suggests significant intermixing and possible intermediate Cu-Al phase formation. Interestingly, the total Cu/Al/Cu thickness of ~46nrn is relatively close to the nominal total. In addition, the interfaces between Py-rich and Cu-rich layers are 118 more planer than the ones between Al-rich and Cu-rich layers, which display considerable waviness. Cu, FeMn Si Substrate 5011 m Figure 5.27 TEM bright field image of an annealed spin-valve with nominal thickness in nanometers Nb(1 5 0)Cu(5)F eMn(8)Py(24)Cu(l 0)Al(30)Cu( l 0)Py(24)Cu(5)Nb(50). 119 Figure 5.28 Phase contrast TEM image of Py/Cu/Al/Cu/Py layers in the annealed spin-valve with x=30nm. Figure 5.29 shows an HRTEM image in the Al/Cu/Py region. Five areas with clearly resolved lattice images were analyzed using FFT. Area 1 in the Al-rich layer shows an AlzCu <100> zone pattern. The entire Cu-rich layer was identified as the [3’ ordered AlCu3 phase. FFTS from this Cu-rich layer show different zone patterns: area 2 shows an AlCu3 <111> zone pattern and area 3 shows an AlCU3 <001> zone pattern. The observations of A12Cu and AlCu3 suggest significant intermixing of Al and Cu have taken place. The Py layers have been found to form metastable bcc structures in the spin- valves. This is illustrated in the FFT analysis from area 4, which shows a Py <111> zone 120 pattern with three-fold symmetry (<111> zones in fee Py have lattice spacings that cannot be resolved with the JEOL 2200FS). However, most of the Py in the spin-valves exhibit fee structure. A twinned grain of fee Py is observed in area 5, with the corresponding FFT shown at the bottom of figure 5.29. It is clear that all the phases have grown on their close packed planes. The lattice spacings measured from different phases are summarized in table 5.4. This table shows that these nano-scaled phases exhibit lattice parameters consistent with the bulk phases. 121 2-20 ‘_ ' '1“ 15‘ um, [our] I=('(' l’_\> <1 10> Figure 5.29 HRTEM image of the Al/Cu/Py layers in the armealed spin-valve with x=30nm. Five different areas in the image are analyzed using FFTS. Areas 1 shows a 0 AlzCu <100> zone pattern, area 2 shows a [3’ AlCu3 <111> zone pattern, area 3 shows a B’ AlCu3 <001> zone pattern, area 4 shows a BCC Py <111> zone pattern, and area 5 shows a FCC Py <111> zone pattern. 122 Table 5.4 Summary of the lattice spacings measured from all the phases in figure 5.29. Theoretical Measured Layer Plane . . spacrng (nm) spacrng (nm) (111) 0.205 0.20 Py (fcc) (200) 0.178 0.18 Py (bcc) (110) 0.205 0.20 (110) 0.205 0.20 AlCu3 (200) 0.145 0.14 (020) 0.302 0.30 AlzCu (002) 0.243 0.24 (022) 0.189 0.19 Complementary XEDS line scans were carried out to investigate the chemical nature in the Al/Cu/Py region. Figure 5.30a shows a STEM bright field image of the Py/Cu/Al/Cu/Py layers in the annealed spin-valve. XEDS intensity profiles of Cu, Ni, Fe, Al and 0 were collected fiom the line shown across the Py/Cu/Al/Cu/Py layers. The line profiles indicate that the Al intensity profile extends through the adjacent Cu layers while Py (Ni and Fe) intensity profiles do not. This observation suggests that there is no obvious intermixing between the Py (Ni and Fe) and Cu region, while significant intermixing taking place between the Cu and Al region. This chemical nature analysis is consistent with the results from the HRTEM analysis. Both HRTEM and XEDS analyses show that intermediate Cu-Al phases were conclusively identified in the Al(30nm) annealed sample. 123 (a) Copper Ka1 , Nickel K81 , Iron Ka1 , Oxygen K31 Figure 5.30 (a) STEM bright field image of the multilayers in an annealed spin- valve. (b) XEDS intensity profiles of Cu, Ni, F e, Al and O are collected from the line across the Py/Cu/Al/Cu/Py layers in the image (a). 124 5.3 Discussion and Conclusions Although the interfacial intensity profiles for the as-sputtered and annealed F(Co, or Py)/Al multilayers suggest some limited intermixing exists, both HRTEM and diffraction studies show no obvious intermediate phase formation. In particular, the annealing treatments do not significantly alter the multilayer structures. The equilibrium phase diagram of the Co—Al system shows that the B2-ordered CoAl phase has a very high melting temperature, which suggests this phase is very stable in terms of volume free energy. However, there is no obvious evidence showing the any intermediate phase formed in the Co/Al multilayers. It is well known that when a material’s dimension decreases to the nanometer scale, the surface and interfacial energy become more significant and their contributions to the total free energy cannot be ignored. Consequently, the absence of any intermediate phase in the Co/Al multilayers suggests the surface energy of the intermediate phase should be high and /or the interfacial energy between the Co and Al, and the CoAl phase should be high, which inhibits nucleation of the B2 phase. However, rapid formation of the A19C02 and B2-ordered CoAl phases were observed in Co(20nm)/Al(50nm) bilayer atom probe tips annealed at 300°C for 5min [47]. This may suggest the surface energy/volume energy balance of these intermetallic phases alters depending on the film thickness. For example, it is lcnown that Co and Cu are insoluble in bulk systems [19]. However, at the nanometer scale, solid solutions of Co and Cu have been observed by X-ray diffraction and calorimetric measurements [68]. In addition, thermodynamic calculations show that the solubility in Co/Cu multilayers varies depending on the layer thickness [69]. Consequently, the phase stability in different layer thickness can give valuable insight for rational design of multilayers structures. 125 HRTEM and XEDS microanalysis of the as-sputtered and annealed Cu/Al multilayers show obvious intermediate phase formation. A12Cll and AlCu3 are formed when the overall stoichiometry is 52.9 at.% Cu in the Cu(8nm)/Al(10nm) multilayers, while AlgCu and Cu are formed when the overall stoichiometry is 70.1 at.% Cu in the Cu(5nm)/Al(3nm) multilayers. For Cu/Al/Cu layers in the spin-valves, evidence of AlzCu and AlCU3 phase formation in the annealed spin-valve with the 30nm Al layer was found, while AlzCu and Cu were observed in the as-sputtered spin-valve with the 10nm Al layer. The overall Al/Cu stoichiometry is 48.4 at.% Cu in the as-sputtered spin-valve with Cu(10nm)/Al(10nm)/Cu(10nm) layers and 73.7 at.% Cu in the annealed spin-valve with Cu(lOnm)/Al(30nm)/Cu(10nm) layers. It is interesting that the intermediate phases formed in these multilayers are similar in some extent if the overall stoichiometries are close. This reinforces the critical concept that phase formation in small system is associated with stoichiometries, volume energies, and interfacial energies. Kung and co-workers [70] have observed both fee and bcc Cu in the CW multilayers. That study found that the variation of the Cu structures depended on the layer thickness between 1.2 and 100 nm with constant volume fiaction of Cu of 50%. For layer thickness above 2.5nm, fcc Cu and bcc Nb formed a Kurdjumov-Saches orientation relationship. However, for layer thickness under 1.2, a slight distortion of the bee Cu was observed. Of critical note, they found that the lattice mismatch play a more important role when the layer thickness decreased below a critical value. No intermediate phases were observed, consistent with the Cu-Nb phase diagram, which shows that these metals are insoluble. Geng and co-workers [71] have observed nonequilibrium bcc Cu structures in Py/Cu spin valves with Cu layer adjacent Nb contact 126 layer. HRTEM and FF T analyses from that study show that in some columnar grains in the spin valve bcc Cu grew on {110} planes and was epitaxial with the Nb contacts. In the case of Cu/Al layers observed in the present study, interfacial energies of the intermediate phases will vary with the overall stoichiometry of the Cu and Al in the multilayers. Not only will the mismatch contribution to the interfacial energy be a function of the stoichiometry of the phase (which in turn can be a function of the overall stoichiometry) due to lattice parameter effects, but the interfacial energy will also be a function of the changes in chemical bonding associated with stoichiometry changes. In summary, no intermediate phase formation is observed in both as-sputtered and annealed Co/Al and Py/Al multilayers. Any interfacial mixing, if it exists, is limited in scale. Large lattice mismatches result in the interfaces tilting significantly to accommodate the strain. Different intermediate phases were observed in the Cu/Al multilayers and Cu/Al regions in the spin-valves in both as-sputtered and annealed conditions. The phases formed vary with the overall stoichiometry of the Cu and Al. Nevertheless, the intermediate phases always grow on their close-packed planes and orient in the direction with the minimum lattice mismatches. 127 Chapter 6 Overall Discussion and Summary This dissertation presents an HRTEM and EELS study of a number of nanoscale structures. The critical issues that have been focused on include crystal structures, stoichiometries, phase stabilities, interfacial epitaxy, interfacial intermixing, and the electronic nature of these systems. It is well known that materials can behave far from their equilibrium states when their dimensions decrease to the nanometer scale. At this scale, the surface and interfacial energies become non-negligible factors in the behavior of materials. The selected systems in this study provide basic insights into the balances in thermodynamic volume fiee energies and interfacial energies that affect nucleation, growth, and stability of materials at the nanoscale. The understanding of the structure and stability of such nanoscale materials are critical to realize rational design and synthesis with desired properties. In the studies of Gd silicide nanostructures, it has been found that both hexagonal GdSiz.x and thet GdSi; can self-assemble on Si(lOO). FFT analysis shows that the lattice parameters of the nanostructure phases do not deviate significantly from the bulk values. In the equilibrium Gd-Si system, the orth/tet GdSiz phase has a high congruent melting temperature, significantly higher than the pertectoid temperature that forms the hexagonal phase (Gd-Si phase diagram shown in figure 2.4), suggesting the orth/tet phase is more stable than the hexagonal GdSiz.x phase in terms of volume free energy. This is consistent with the observation in this study that the orth/tet Gd silicide thin films form from the annealing of alternating Gd/Si layers. The observation that Gd silicide nanostructures with the orth/tet structure form after extended post deposition annealing also suggests that this phase has a lower volume free energy. Despite the orth/tet phase 128 being the lower energy phase, the hexagonal phase is found to form in the early stages of growth. This is consistent with the interfacial energy dominating during the initial growth stage, which is far from equilibrium. While the isotropic lattice mismatch of the hexagonal phase with Si(lOO) is deemed responsible for the formation of nanowires, the overall surface energy of the hexagonal/Si(l 00) interface associated with this isotropic mismatch must be lower than that of the orthorhombic/Si( 1 00) interface in order to overcome the larger volume free energy of the hexagonal phase. The observation of overgrowth of the orth/tet phase over the hexagonal phase shows that the hexagonal phase is energetically favorable at fine scale and can act as a buffer layer between the orth/tet phase and Si substrate. The reason for this is that the hexagonal phase has much lower 2D lattice mismatch than the orth/tet phase with Si substrate which in turn results in a lower interfacial energy. A similar balance between volume fi'ee energy and interfacial energy is found in the Tm-Si system, as discussed in section 4.2.2. ELS spectroscopy of Gd, Gd oxide and Gd silicide failed to provide insight into the differences in chemical environment of the Gd in these materials, other than to verify that Gd is indeed trivalent in all three cases. In this respect, the three materials selected were a poor test case to explore the capabilities of a newly installed ELS system to provide detailed chemical bonding information. Nevertheless, the data presented here are a reference point for future studies. On the other hand, the sensitivity of the ELS spectra to plasmon related features provides useful information on metallic multilayer systems. In the case of metallic multilayers, the interfacial energy/volume energy balance also determines the resulting nanoscale structures. This is revealed in the different systems by varying types of phase equilibria. 129 The Co-Al binary phase diagram indicates four different intermetallic compounds that might be expected to form by interdiffusion of the pure Al and Co layers at the growth and annealing temperature. Of these, the B2-ordered CoAl phase has a high congruent melting temperature of ~1 648 °C , much higher than the melting temperatures of Al and Co at 660 °C and 1494 °C respectively [19], suggesting this CoAl phase should be very stable. Surprisingly, the HRTEM and diffraction studies show no evidence of formation of this B2 phase, or any other intermetallic phase formation. This lack of phase formation persists with annealing. This absence of any intermediate phase formation in the Co/Al multilayers suggests the interfacial energy between the Co and/or Al and the potential intermediate phases is high, which inhibits nucleation of the intermediate phases. This high interfacial energy may be a result of significant mismatch between the base components and the intermetallic phases. Indeed, the bee derivative B2 structure will not form a simple coherent interface with either fcc Al or hcp Co, and might be expected to require formation of complex variation of a typical fcc/bcc interface such as a Kurdjumov-Sachs or Nishiyama-Wassermann relationship. Likewise, the other intermetallic phases in this system, monoclinic A19C02, monoclinic A13Co4, and hexagonal A15C02 (listed in table 2.7) would be expected to form complex interfaces with Al and Co as these phases have relatively complex crystal structures. Similarly, the equilibrium binary (F e-Al, Ni-Al) [l9] and ternary (Ni-Fe-Al) [72] phase diagrams show that a number of intermetallic phases including B2 F eAl (with melting temperature of ~1310 °C), B2 AlNi (with melting temperature of ~1 638 °C), L12 AlNi3 (with melting temperature of ~1 395 °C) [19], monoclinic FeNiA19 (with melting temperature of ~809 °C), and F e3NiAllo (with a poorly defined crystal structure and a 130 melting temperature of ~1 050 °C) [72] might be stable and expected to form in the PyAl multilayers. However, the relatively complicated crystal structures of these intermetallic phases would be expected to form more complex interfaces with high surface energies. Consequently, the complex interfaces between the Ni (and/or Fe) and Al intermediate phases and the Py and Al might be expected to inhibit nucleation of the intermediate phases. An important objective of this study has been to understand the structure of the metallic multilayers in an effort to correlate their observed magnetoresistance behaviors with the nanoscale phase and interface structure. The idealized design of multilayers for devices requires both a large value of the enhanced specific resistance (ZAR; , M) and a large scattering asymmetry (y) (defined in section 2.3.3). For the case of the F(Co, Py)/Al multilayers, the measurements show similarly large values of 2AR* ~10 fflmz, but small values of y _<_ 0.1 for both systems, which suggests they are not completely idealized for devices [46]. It is also observed that these parameters vary with aging and annealing, specifically with the AR increasing by 5-10% after 6-11 months and further by 2-7 % on annealing at 180°C for less than 5 min, suggesting subtle changes in the structures [46]. However, the TEM observation in this study found no obvious intermediate phase formation and interfacial structure change. Consequently, the abnormal MR behavior changes observed in the F(Co or Py)/Al multilayers cannot be attributed to changes in interfacial structure or intermediate phase formation. Other changes, such as (1) variations in the stacking fault densities in the multilayers, and (2) changes in electronic nature of the transition metals (Co, Ni and Fe) at the interfacial region caused by interfacial ordering changes, may affect the multilayer MR behavior. In 131 order to attribute the observed changes in MR behavior to the changes in stacking fault density, a thorough statistical analysis measuring the stacking fault probability across the multilayers is necessary. Since XRD is more effective than TEM in determining the stacking fault densities across larger volumes, further analysis combining extensive TEM measurements with x-ray diffraction (XRD) may be useful for understanding the MR changes with aging and annealing. The electronic nature of the transition metals (Co, Ni and Fe) at the interfacial region can be experimentally investigated by studying the near- edge fine structures of these elements for chemical shifts and relative intensities changes of the L-edges. However, such detailed information requires higher resolution EELS capabilities which are not available at MSU. Although the Cu/Al multilayers were not originally synthesized for GMR devices, these multilayers were produced to obtain fundamental data on the CPP interface specific resistance (area A times resistance R). Such data is critical for interpreting the large AR* in F/Al multilayers. However, the TEM results indicate that correlating the data of Cu/Al multilayers to F/Al multilayers is not straightforward because the HRTEM and XEDS microanalyses of both the as-sputtered and annealed Cu/Al multilayers show obvious intermediate phase formation. The formation of these different intermediate phases is quite interesting and can again be related to the balance between the volume free energies and surface and/or interfacial free energies. Based on the equilibrium phase diagram of Cu and Al (figure 2.17), one might expect that 112 AlCu (monoclinic) and C2 (poorly defined) phases (see table 2.8 for the crystal structures in the Al/Cu system) will form when the overall stoichiometry is 52.9 at.% Cu in the Cu(8nm)/Al(10nm) multilayers (based on a simple 132 lever rule calculation, the resulting structure would be expected to be 47.3% 112 and 52.7% C2). Similarly, 69% 71 A14Cu9 (complex cubic) and 31% a2 (poorly defined) will form when the overall stoichiometry is 70.1 at.% Cu in the Cu(5nm)/Al(3nm) multilayers. However, our results show that AlzCu (tetragonal) and AlCu; (disordered bcc) are formed when the overall stoichiometry is 52.9 at.% Cu (Cu(8nm)/Al(10nm) multilayers), while AlzCu and Cu are formed when the overall stoichiometry is 70.1 at.% Cu (Cu(5nm)/Al(3nm) multilayers). The observed Cu-Al intermetallic phases were grown with preferred orientation relationships and relatively small lattice mismatches, suggesting the interfacial energy is small. Other intermetallic phases such as monoclinic AlCu, hexagonal A13Cu4, and cubic A14CU9 have not been observed in the multilayers because more complex interfaces are expected to form. Consequently, the high surface energy prohibits the nucleation of these phases. In addition, different phases are observed in Cu/Al multilayers with different layer thickness and overall stoichiometry suggesting that the balance between thermodynamic volume fi'ee energies and interfacial energies of the phases depends on the layer thickness and overall stoichiometry. The observed intermediate phases are expected to have significantly different ARs than the base Cu and Al. Table 6.1 lists the electrical resistivity data of pure Al, pure Cu (fee), and Cu/Al alloys Al(67%)/Cu(33%) and Al(25%)/Cu(75%) that have stoichiometries close to AlzCu and A1Cll3 phases respectively [73]. Nevertheless, it should be noted that the bee solid solution AlCu3 is not a stable phase at room temperature and its electrical resistivity is not practically measurable. Consequently, the alloy Al(25%)/Cu(75%) shown in the table should only be considered as an approximation for the resistivity of this phase. As both the 8/10 and 5/3 Cu/Al 133 multilayers contain the tetragonal A12Cu phase in similar proportions (roughly 2:1- Al- rich phase to Cu-rich phase), it would be expected that differences in the Cu-rich phase will have a strong influence on the measured AR. When the sputtered Al is relatively thick, the AlCu; bcc solid solution is observed with the tetragonal AlzCu phase: Since AlCu3 stoichiometry has large resistivity, the overall AR will be large. When the sputtered Al is relatively thin, fcc Cu is observed: Since Cu has a small resistivity, the overall AR will be low. Since the intermediate phases formed in both Al/Cu multilayers and the Cu/Al/Cu region of the Py/Cu/Al based spin-valves are similar to some extent if the overall stoichiometries are close (discussed in section 5.3), the relationship between the transport and possible intermediate phase formation should be considered when designing GMR devices containing Cu/Al layers [48]. Table 6.1 Electrical resistivity data of the phases in the Cu/Al multilayers, after [73]. Electrical resistivit Alloy (LI 9 cm) Y Al 2.49 Cu (fcc) 1.54 Al(67%)Cu(33%) ~6 Al(25%)Cu(7 5%) ~20 In summary, a number of nanoscale structured electronic materials have been studied using a broad range of microanalytical techniques, including HRTEM, complimented with XEDS, EELS, and STEM. Characterization of the crystal structures, interfacial epitaxy, interfacial chemical nature, and electronic nature has lead to a 134 comprehensive understanding of the balance between volume free energy and interfacial energy determining the phase formation and phase stability on the nanometer scale. These fundamental structural studies are correlated with their physical properties in an effort to realize rational design and applications of these nanoscale electronic materials. The possibility of intermediate phase formation, in equilibrium and nonequilibium states, should be taken into account when designing electronic materials structured at the nanometer scale. 135 Chapter 7 Suggestions for Future Work 7.1 Patterned Growth Studies of Silicide Nanostructures The epitaxial growth of rare-earth (RE) metal silicides can result in nanowires and nanoislands self-assembling on Si(001) due to the mismatch between the silicides and the Si(001) surface. However, these nanostructures tend to grow in random locations (and unless a vicinal substrate is used, in two directions) instead of at specific locations, since the nucleation of the silicides on the Si substrate is random. In order to prompt these nanostructures to grow in desired patterns and locations, one would have to imagine a way to design patterns for controlled nucleation first. There are advanced techniques available, such as focused ion beam (FIB) nanopatterning, which may realize nucleation in a designed pattern. FIB (with a high resolution of ~2nm) can be used to mill trenches or holes in the substrate in a pattern before the metals are deposited, and in this way, the nucleation might be energetically favorable at these trenches or holes. In addition, one can also take advantage of FIB for TEM sample preparation. STM and TEM study (plan- view and cross-sectional) may be more closely correlated by making TEM samples of specifically identified nanowires and structures using FIB. In particular, specimens of both orientations (plan-view and cross-sectional) can be obtained from the same growth sample, which is not possible with the sandwich technique described in this thesis. In this way, the nanowire and island morphologies of the silicide nanostructures can be more clearly correlated with the crystal structures. Specific features, such as nanostructure junctions, can be studied in detail. Consequently, the nucleation and growth mechanism of the nanostructures can be more clearly understood in the 136 combination of plan-view and cross-sectional TEM. Moreoever, the yield of usable samples from a small number of growth samples would greatly increase. 7.2 High-resolution EELS Analysis Using Monochromated TEM Monochromated transmission electron microscopes are now commercially available and have been reported to have full-width at half-maximum (FWHM) energy resolutions of 0.1 - 0.25 eV with a cold field-emission electron gun (CFEG) [74]. One of the advantages of the monochromator is that it allows much higher resolution EELS analysis to be carried out, which can be used to more thoroughly explore the band structure of electronic materials. Although X-ray absorption has a better energy resolution (0.03 eV), its spatial resolution (50 um) is not suitable for investigation of the electronic nature of individual nanoscale structures. Lazar and co-workers [75] have studied the near-edge fine structure of Ga L2,; and N ionization edges in zinc—blend GaN and wurtzite GaN using high-resolution EELS (energy resolution: 0.2 eV) in a monochromated TEM. The experimental spectra were compared with calculations based on the density functional theory using the Wien2k code. Different electronic behaviors in the two crystal structures were correlated with the density of states in different bonding environments. In a similar manner, by studying the near-edge fine structures of RE silicide nanostructures with different phases using HR- EELS, the electronic nature of the different phases and at the interfaces could be more clearly understood. 137 7 .3 EELS Quantification Analysis in Magnetic Multilayers The abnormal MR behavior observed in the F (Co or Py)/Al multilayers can not be attributed to the intermediate phase formation based on the observation in this thesis. However, the electronic nature of the transition metals (Co, Ni and Fe) can be affected by the changes in bonding in the interfacial region, which in turn may affect the entire multilayer MR behavior. This information can be experimentally investigated by studying the near-edge fine structures of the atoms, such as the chemical shifts and relative intensities of the L-edges (white lines in the Co, Ni, and Fe). Similar near-edge fine structure characterization could also be carried out on Cu/AJ multilayers. Kung and co-workers [70] have observed a 2 eV shift in the Cu L3 edge in fee and bcc Cu using EELS characterization. This result suggests that bcc Cu has a lower Fermi energy than fcc Cu. Using this approach to study the Cu/Al multilayers, the difference of the electronic structure of Cu in the fcc Cu and the intermediate Cu/Al phases (AlzCu, ordered and disordered AlCu3) could be determined. Thus, using a monochromated TEM with high energy resolution EELS to study the near-edge fine structure of the transitional metals in the multilayers using spatially resolved EELS may give some clues of the origin of the abnormal MR behaviors. 138 REFERENCES . J. Solomon, Chemical Industry R8D Roadmap for Nanomaterials by Design: From Fundamentals to Function (2003); http://www.chemicalvision2020.org/nanomaterialsroadmap.html . D. B. Williams and C. B. Carter, Transmission Electron Microscopy: A Textbook for Materials Science, ISBN-10: 030645324X (Plenum, New York, 1996). . R. F. Egerton, Electron Energy-Loss Spectroscopy in the Electron Microscope, 2'” ed., ISBN-10: 0306452235 (Plenum, New York, 1996). . R. D. Leapman, L. A. Grunes, and P. L. Fejes, Study of the L23 edges in the 3d transition metals and their oxides by electron-energy-loss spectroscopy with comparisons to theory, Phys. Rev. B, 26, p.614-635 (1982). . R.F. Egerton and M. J. Whelan, Electron energy-loss spectra of diamond, graphite and amorphous carbon, J. Electron Spectrosc., 3, p.232-236 (1974). . D. Pine, Elementary Excitations in Solids, ASIN: BOOOIG3ZO8 (W .A. Benjamin, New York, 1963). . H. Kurata and C. Colliex, Electron-energy-loss core-edge structures in manganese oxides, Phys. Rev. B, 48, p.2102-2108 (1993). . C. Preinesberger, S. Vandré, T. Kalka, and M. Dahne-Prietch, Formation of dysprosium silicide wires on Si(001), J. Phys. D, 31, p.L43-L45 (1998). . Y. Chen, D.A.A. Ohlberg, G. Medeiros-Ribeiro, Y.A. Chang, and RS. Williams, Self-assembled growth of epitaxial erbium disilicide nanowires on silicon (001), Appl. Phys. Lett., 76, p.4004-4006 (2000). 10. J. Nogami, B.Z. Liu, M.V. Katkov, C. Ohbuchi, and NO. Birge, Self-assembled rare-earth silicide nanowires on Si(001), Phys. Rev. B, 63, p.233305-8 (2001). 139 11. Y. Chen, D. A.A.Ohlberg, and R. S. Williams, Nanowires of four epitaxial hexagonal silicides grown on Si(001), J. App. Phys., 91, p.3213-321 8 (2002). 12. C. Ohbuchi and J. Nogami, Holmium growth on Si(001): Surface reconstructions and nanowire formation, Phys. Rev. B, 66, p.165323 (2002). 13. R. Ragan, Yong Chen, D. A.A.Ohlberg, G. Medeiros-Ribeiro, and RS. Williams, Ordered arrays of rare-earth silicide nanowires on Si (001 ), Journal of Crystal Growth, 251, p.657-661 (2003). 14. D. Lee and S. Kim, Formation of hexagonal Gd disilicide nanowires on Si(100), Applied Physics Letters, 82(16), p.2619-2621 (2003). 15. M. Kuzmin, P. Laukkanen, R.E. Perala, R.-L. Vaara, and 1.]. Vayrynen, Formation of ytterbium silicide nanowires on Si(001), Applied Surface Science, 222, 394 (2004). 16. M.V. Katkov and J. Nogami. Yttrium growth on Si(001). in American Physical Society (APS) March Meeting (2002). 17. J. Zhang, M. A. Crimp, Y. Cui, and J. Nogami, Self-assembled Tm silicide nanowires grown on Si(001) studied by ST M and T EM, in preparation. 18. A. Vantomme, M.F. Wu, S. Hogg, U. Wahl, W. Deweerd, H. Pattyn, G. Langouche, S. J in, and H. Bender, Stabilisation and phase transformation of hexagonal rare-earth silicides on Si(I I I), Nucl. Instrum. Meth. Phys. Res. B, 147, p.261-266 (1999). 19. TB. Massalski, H. Okamoto, P.R. Subramanian, and L. Kacprzak, Binary alloy phase diagrams. Vol. 2. 1990, Metals Park, OH: ASM International. 20. G. Ye, Thesis of Rare-earth Silicide Nanostructure Epitaxial Growth and Crystallography, Ph.D. dissertation, Michigan State University, East lansing, Michigan (2006). 140 21. G. Ye, J. Nogami, and M.A. Crimp, Dysprosium disilicide nanostructures on Si(001) studied by scanning tunneling microscopy and transmission electron microscopy, Thin Solid Films, 497(1-2), p.48-52 (2006). 22. K.N. Tu, R.D. Thompson, and B.Y. Tsaur, Low Schottky barrier of rare-earth silicide on n-Si, Applied Physics Letters, 38(8), p.626-628 (1981). 23. H.W. Yeom, Y.K. Kim, E.Y. Lee, K.-D. Ryang, and P.G. Kang, Robust One- Dimensional Metallic Band Structure of Silicide Nanowires, Physics Review Letters, 95, p.205504 (2005). 24. C. Preinesberger, G. Pruskil, S.K. Becker, M. Dahne, D.V. Vyalikh, S.L. Molodtsov, C. Laubschat, and F. Schiller, Structure and electronic properties of dysprosium-silicide nanowires on vicinal Si(001), Applied Physics Letters, 87(8), p.083107 (2005). 25. K Suenaga, S. Iijima, H. Kato and H. Shinohara, F ine-structure analysis of Gd M45 near-edge EELS on the valence state of Gd@C82 microcrystals, Phys. Rev. B, 62, p. 1627-1630 (2000). 26. T. Manoubi, C. Colliex and P. Rez, Quantitative electron energy loss spectroscopy on M 45 edges in rare earth oxides, J. Electr. Spectr. and Rel. Phen., 50, p.1-18 (1990). 27. C. C. Ahn, ed., Transmission electron energy loss spectrometry in materials science and the EELSATLAS, 2'“ ed., ISBN-10: 3527405658 (Wiley, Weinheim, 2004). 28. A. Travlos, P. Aloupogiannis, E. Rokofyllou, C. Papastaikoudis, A. Traverse, G. Weber, Growth, characterization and electrical properties of gadolinium silicide thin layers, Philosophical Magazine B, 67, p.485-495 (1993). 29. R. G. Long, M. C. Bost, and J. E. Mahan, Metallic behavior of lanthanum disilicide, Applied Physics Letters, 53, p.1272-1273 (1988). 30. A. Barthélémy, A. Fert, J-P. Contour, M. Bowen, V. Cros, J. M. De Teresa, A. Hamzic, J. C. Faini, J. M. George, J. Grollier, F. Montaigne, F. Pailloux, F. 141 31. 32. 33. 34. 35. 36. 37. 38. 39. Petroff, and C. Vouille, Magnetoresistance and spin electronics, Journal of Magnetism and Magnetic Materials, 242-245, p.68-72 (2002). G. Binash, P. Griinberg, F. Saurenbach, and W. Zinn, Enhanced magnetoresistance in layered magnetic structures with antiferromagnetic interlayer exchange, Phys. Rev. B, 39, p.4828-483O (1989). M. Baibich, J .M. Broto, A. Fert, F.N. Guyen Van Dau, F. Petroff, P. Etienne, G. Creuzet and A. F riederich, Giant Magnetoresistance of (001 )F e/(001 )Cr Magnetic Superlattices, Phys. Rev. Lett., 61, p.2472-2475 (1988). R. Loloee, Growth and Characterization of Epitaxial Films and Magnetic Multilayers Grown by DC Sputtering, Ph.D. dissertation, Michigan State University, East lansing, Michigan (2000). J. Bass and W. P. Pratt Jr., Current-perpendicular (CPP) magnetoresistance in magnetic metallic multilayers, Journal of Magnetism and Magnetic Materials, 200, p.274-289 (1999). B. Dieny, V.S. Speriosu, S.S.P. Parkin, B.A. Gurney, D.R. Wilhoit and D. Mauri Giant magnetoresistive in soft ferromagnetic multilayers, Phys. Rev. B, 43, p.1297-13OO (1991). V.S. Speriosu, B. Dieny, J .-P. Nozieres, B.A. Gurney, P. Humbert, H. Lefakis, P. Baumgart and D. R. Wilhoit, Magnetoresistive properties of spin valves, Digest of the INTERMAG’92, Conference, paper CA-04 (1992). S. Urazhdin, N. O. Birge, W. P. Pratt, Jr., and J. Bass, Switching current versus magnetoresistance in magnetic multilayer nanopillars, Appl. Phys. Lett., 84, p.1516-1518 (1995). N. Theodoropoulou, A. Sharma, R. Loloee, W.P. Pratt, Jr., J. Bass, A. Fert, and H, J affi'es, Interface Specific Resistance and Scattering Asymmetry of Permalloy/Al, J. Appl. Phys., 99, 08G502 (2006). K. Le Dang, P. Veillet, E. Velu, S.S.P. Parkin and C. Chappert, Influence of crystal structure on the magnetoresistance of Co/Cu multilayers, Appl. Phys. Lett., 63, p.108-110 (1993). 142 40. 41. 42. 43. 45. E. E. Fullerton, D. M. Kelly, J. Guirnpel, I.K. Schuller, and Y. Bruynseraede, Roughness and giant magnetoresistance in F e/Cr superlattices, Phys. Rev. Lett., 68, p.859-862 (1992). V. S. Speriosu, J. P. Nozieres, B. A. Gurney, B. Dieny, T. C. Huang and H. Lefakis, Role of interfacial mixing in giant magnetoresistance, Phys. Rev. B, 47, p.11579-11582 (1993). J. Kudmovsky, V. Drchal, I. Turek, C. Blaas, P. Weinberger, and P. Bruno, The influence of defects on electrical transport in magnetic multilayers, J. Mater., 52(7), p.29 (2000). A. G. F. Garcia, 1. N. Krivorotov, P. M. Braganca, D. C. Ralph, and A. R. Buhnnan, Bull. Am. Phys. Soc., 50, p.237 (2005). . N. Theodoropoulou, A. Sharma T. Haillard, R. Loloee, W.P. Pratt Jr., J. Bass, J. Zhang and M.A. Crimp, Specific Resistance, Scattering Asymmetry, and some Thermal Instability, of Co/Al, F e/Al, and C091Fe9/Al Interfaces, IEEE Transactions on Magnetics, 43(6), p.2860-2862 (2007). A. Zambano K. Eid, R. Loloee, W.P. Pratt Jr., and J. Bass, Interfacial Properties of F e/Cr Multilayers n the Current-Perpendicular-to-Plane Geometry, Journal of Magnetism and Magnetic Materials, 253, p.51-55 (2002). 46. A. Sharma, N. Theodoropoulou, T. Haillard, R. Loloee, W. P. Pratt. Jr., J. Bass, J. Zhang and M. A. Crimp, Current-perpendicular to plane (CPP) magnetoresistance of Ferromagnetic (F )/Al interfaces (F = Py, Co, Fe and C091Fe9) and structural studies of Co/Al and Py/Al. (in preparation) 47. V. Vovk, G. Schmitz, and R. Kirchheim, Nucleation of product phase in reactive diflusion ofAl/Co, Phys. Rev. B, 69, p.104102 (2004). 48. A. Shanna, N. Theodoropoulou, R. Loloee, W. P. Pratt. Jr., J. Bass, J. Zhang and M. A. Crimp, Current-perpendicular to plane (CPP) Resistance of Cu/Al Interfaces. (in preparation) 143 49. J .C. Bravman and R. Sinclair, The Preparation of Cross—section Specimens for Transmission Electron Microscopy, Journal of Electron Microscopy Technique, 1, p.52-61 (1984). 50. RA. Stadelmann, EMS - a software package for electron difi'action analysis and HREM image simulation in materials science, Ultramicroscopy, 21, p.131-145 (1987) 51. T. Manoubi, C. Colliex and P. Rez, Quantitative electron energy loss spectroscopy an M4 5 edges in rare earth oxides, J. Electr. Spectr. and Rel. Phen., 50, p.1-18 (1990). 52. H. Kurata and C. Colliex, Electron-energy—loss core-edge structures in manganese oxides, Phys. Rev. B, 48, p.2102-2108 (1993). 53. I. Gerocs, G. Molnar, E. Jaroli, E. Zsoldos, G. Peto, J. Gyulai and E. Bugiel, Epitaxy of orthorhombic gadolinium disilicide on <100> silicon, Applied Physics Letters, 51(25), p.2144-2145 (1987). 54. G. Molnar, I. Gerocs, G. Peto, E. Zsoldos, J. Gyulai and E. Bugiel, Epitaxy of GdSig 1.7 on <111> Si by solid phase reaction, Applied Physics Letters, 58(3), p.249-250 (1991). 55. D. Lee and S. Kim, Formation of hexagonal Gd disilicide nanowires on Si(100), Applied Physics Letters, 82(16), p.2619-2621 (2003). 56. G. Ye, J. Nogami, and M.A. Crimp, Crystallographic study of self-assembled dysprosium silicide nanostructures on Si(001), Phys. Rev. B, 74, p.033104 (2006). 57. A.J. Nelson, J .J . Adams, and KL Schaffers, Electronic structure of gadolinium calcium oxoborate, Applied Surface Science, 252(5), p.1228-1232 (2005). 58. E. Puppin, J. Nogami, C. Carbone, Z. X. Shen, I. Lindau , B. B. Pate, I. Abbati and L. Braicovich, A photoemission study of the Si(111)/Gd interface: A comparison with the bulk silicides, Journal of Vacuum Science & Technology B, 5, p.1083-1086 (1987). 144 59. A. V. Fedorov '1, E. Arenholz, K. Starke, E. Navas, L. Baumgarten, C. Laubschat, and G. Kaindl, Surface Shifts of 4f Electron-Addition and Electron-Removal States in Gd(0001), , Phys. Rev. Lett. 73, p.601-604 (1994). 60. B. T. Thole, G. van der Laan, J. C. Fuggle, G. A. Sawatzky, R. C. Kamatak and J .-M. Esteva, 3d x-ray-absorption lines and the 3d94f+1 multiplets of the lanthanides, Phys. Rev. B, 32 (8), p.5107-5118 (1985). 61. Z. Hu, G. Kaindl, B.G. Muller, Core-level spectroscopy of the tetravalent lanthanide compounds M 3LnF 7 (with M = Cs, Rb; LN = Ce-Pr,Nd,Tb,Dy), J. Alloys Compd., 246, p.177-185 (1997). 62. A. Iandelli, A. Palenzona and G. L. Olcese, Valence fluctuations oy iterbium in silicon-rich compounds, J. Less- Common Metals, 64, p.213-220 (1979) 63. J. Sugar, Potential-Barrier Efl'ects in Photoabsorption. II. Interpretation of Photoabsorption Resonances in Lanthanide Metals at the 4d-Electron Threshold, Phys. Rev. B, 5, p.1785-l792 (1972) 64. Ch. Rath, J. E. Prieto, S. Muller, R. Miranda, and K. Heinz, hcp-to-fcc stacking switch in thin cobalt films induced by Cu capping, Phys. Rev. B, 55, p.10791- 10799 (1997). 65. A F. Guillermet, Critical evaluation of the thermodynamic properties of cobalt, Int. J. Thermophys., 8, p.481-510 (1987). 66. M. J. Mills, Determination of compositional ordering at grain boundaries in boron-doped Ni 3A1, Scripta Met, 23, 2061 (1989). 67. H-J. H. Kung, The Structure and Chemistry of Grain Boundaries in Nickel Alumnide, Ph.D. dissertation, Cornell University, Ithaca, New. York (1991). 68. C. Gente, M. Oehring, R. Bormann, Formation of thermodynamically unstable solid solutions in the Cu-Co system by mechanical alloying, Phys. Rev. B, 48, p.13244-l3252 (1993). 69. W. Cao, J. Zhu, Y. Yang, W.A. Oates, Y.A. Chang, Thermodynamic stability of Co/Cu multilayered nanostructures, Scripta Materialia, 53, p.1379-1382 (2005). 145 70. 71. 72. 73. 74. 75. H. Kung, Y-C. Lu, A. J. Griffin, Jr., M. Nastasi, T. E. Mitchell, and J. D. Embury, Observation of body centered cubic Cu in Cu/Nb nanolayered composites, Appl. Phys. Lett., 71, p.2103-2105 (1997). H. Geng, J .W. Heckrnan, W.P. Pratt Jr., J. Bass, F.J. Espinosa, S.D. Conradson, D. Lederman, and M.A. Crimp, Occasional ”long-range" nonequilibrium body- centered-cubic structures in NiFe/Cu spin valves, J. Appl. Phys., 86, 4166 (1999). L. Eleno, K. F risk, and A. Schneider, Assessment of the F e-Ni-Al system, Intermetallics, 14(10-11), p.1276-1290 (2006). K. Schré'rder, CRC Handbook of Electrical Resistivities of Binary Metallic Alloys, ISBN-10: 0849335205, p.79 (CRC Press, Inc. 1983). K. Kimoto, G. Kothleitner, W. Grogger, Y. Matsui, F. Hofer, Advantages of a monochromator for bandgap measurements using electron energy-loss spectroscopy, Micron,. 36(2), p.185-189 (2005). S. Lazar, C. Hébert and H. W. Zandbergen, Investigation of hexagonal and cubic GaN by high-resolution electron energy-loss spectroscopy and density functional theory, Ultramicroscopy, 98, p.249-257 (2004). 146 APPENDIX A Detailed TEM Sample Description . Gd oxide powder (in figure 4.1and 4.2): micron sized Gd203 powder (from Alfa Aesar) were grounded to smaller size. . Sputtered Gd/Si layers (in figure 4.3): Si 43 nm/ Gd 75nm/ Si 85 nm/ Gd 75 nm/ Si 43nm layers were deposited in a DC-magnetron triode sputtering system at room temperature. . Bulk GdSi; layers (in figure 4.4): the alternating Gd/Si layers (2) were annealed in a high vacuum of 5><1045 torr at 900 °C for 1 hour to form silicides. . GdSiz nanostructures (STM reference # 041105) (in figure 4.5, 4.6, 4.7, and 4.8): the nanostructures were grown on a Si substrate heated to 650°C with a metal coverage of 1.5ML. This was followed by post-deposition annealing for 20 min at the growth temperature. . Tm silicide nanostructure sample Tm-A (STM reference # Tm060619) (in figure 4.9 and 4.13) was grown with 2.8 ML metal coverage without post- deposition annealing. . Tm silicide nanostructure sample Tm-B (STM reference # Tm082006) (in figure 4.10 and 4.12) grown with 1 ML metal coverage with a post-deposition annealing at 600°C for 30 min. . Tm silicide nanostructure sample Tm-C (STM reference # Tm070215) (in figure 4.14, 4.15, and 4.16) was grown under very similar conditions to Tm-B: 1 ML Tm deposited at 600°C with 20 minutes post annealing. However, this sample was also capped with a layer of amorphous Can. 147 10. ll. 12. l3. 14. 15. As-sputtered Co/Al multilayers (GMR reference 1693-4b) (in figure 5.1, 5.2, 5.3, and 5.6): alternating Co and Al layers with 20 nm Co and 10 nm Al were sputtered on Si(lOO) substrates. Annealed Co/Al multilayers (GMR reference l693-4b) (in figure 5.5): as- sputtered multilayers were annealed for less than 5 min at 180°C after sputtering in the same chamber. As-sputtered Py/Al multilayers (GMR reference 1693-4a) (in figure 5.10): alternating Py and Al layers with 20 nm Py and 10 nm Al were sputtered on Si(lOO) substrates. Annealed Py/Al multilayers (GMR reference 1693-5b) (in figure 5.10, 5.11): as-sputtered layers were annealed for less than 5 min at 180°C after sputtering in the same chamber. As-sputtered Cu(8nm)/Al(10nm) multilayers (GMR reference 1669-3b) (in figure 5.14 and5.15) Annealed Cu(8nm)/Al(10nm) multilayers (GMR reference l669-3b) (in figure 5.17 and5 .18): as-sputtered layers were annealed for less than 5 min at 180°C after sputtering in the same chamber. As—sputtered Cu(5nm)/Al(3nm) multilayers (GMR reference 1669-5b) (in figure 5.19 and 5.20) As-sputtered Nb(1 50)Cu(5)FeMn(8)Py(24)Cu(l0)Al(10)Cu(10)Py(24)Cu(5)Nb(50) spin- valve (GMR reference l689-3a) (in figure 5.22, 5.23, 5.24, 5.25, and 5.26) 148 1 6. Annealed Nb(1 50)Cu(5)FeMn(8)Py(24)Cu(10)Al(30)Cu(]0)Py(24)Cu(5)Nb(50) spin- valve (GMR reference 1689-6a) (in figure 5 .27, 5 .28, 5.29, and 5.30): as- sputtered layers were annealed for less than 5 min at 180°C after sputtering in the same chamber. *‘149 IIi]ll]lli][iil][ilii]lwill]I