in... 1‘... .II . r ...\l|1vu~ut‘ .I‘U‘IIIx thUIIlI . unsunhpywbnfihwhmm. u... . . A . ”Eu Ilfinvuo .1 knitwurutll I" LI. Innocfou . ,n. .0 IIIot.'Ip| 75.1%}.4 .OHU‘unt ll — v 1., «921v» aw. .m.fmm.uummmi...mvn.M-Wfiwl~ I via... .54: .m l A ‘ . . M. A: J. :J . . z. .I .38 (It. I. n I 31‘!) HI: I. 1. Lou . . 0.; . .15“... .quIhtV mm...» fun. . 4 J. . . o! y t _ . . ... I ‘0. ~ ‘ I . :- Ivkrrfi‘us . . . . , .. . I .. .J I LL PI... :Vwflflwflufih . . I .. . , . I. . . ...L...!:;.I.1 “43:81".qu .I I v I .. «A‘L\ 9?. 'IO‘ WWW, z. :3 3g“. _. 13,2 9,. , N . :93“ I 3 o u 3 v . . . H .1 .Ii. , . . , . .23 A. La .. . 5.: .2 . . . . A . :n a a A . . . . . . v rbfllv .. . . I , I u . I 96 I .. . , ... . , .. n. «I 53W... . . .«Imflwlim .zv . . . a n 1.. . ..\. .11.». . Cur! L... {sitessI I. Hat. .. . . I . v 3.“ :hfi UAR . I Ial . , :1 t 19‘ v§!):.rv.1....r|sigu . .. . . 53?... hung. 2...! . . . .2. I . more”. a»... a. Petra : . . 3.59... I . mafia..h .. . t.» e v... : , . ,I . .. . . A) \ ‘ Vt .. ~ . , . .. .. .1 . up . v . . Ibufiaadc . . ,. . . . fi.....?..w...,... , . uni. ‘ .I 5., VI . .. .. . I :1: . ,. . I .wa:r Linwsr . ._= I : a . .. .,g . Iaazz - , ; 1...“... . .. q I... .3... 7.2.... I. , . , . ,. . . . . . .. .. f“... 3.1. I .1 1.53.3.3 é. . Idtrhouho . Man. a .1” v.1.(nflléu. .2Ilwuntbflé . J . I . . .h . . . a » .. .I _ Vina}.- i~.IbI nI‘v his. . ‘3‘: ~ 1.2- .. \_V. .2 . I 3 b5 1'. .é‘ . . «15 1 \ . I . . . . I . nuprl. “6|. “é aw 0| ‘3 033%.. .uowv... . p v Ira...“ j. 5 .123. , . . k . lug-..» a: I. on Iltufi. . €399.35 u, I. Z .... r azufl I» . 1 . II \.I:( P: . . III. . 0.... I... fair .......B. .,... Ina , I _ . . 53:355.. . . .. . . I . . . . . .. . fi..I.J.\....mbn:_.. “human”! .6. .. . . I . .. . Hum}... I....I....:.II....I.....I..:I|I. bu! . . . . I . I I In III-00.-....ID I! . II... n I 1 11.1.53; 3 .v .1 .Ic ,. .‘I. i.... PEI 1!. .o v v t Z... l .V 52%”... a. .. .....kfl2...rm.d1.+ao...3. (4359...: l. c :52: v 2... no. .. i $3.2. I . _ 5b... Ina . . «1.4.33? 1 . . . .1: 1 .. I . .. . . , .. . 5. (It. I :uu . .u . . A. . u I 35 HAIOMHOV$L - , . V . . . ‘ u . _ 1 Hr .Iilflfl :3 J \I \\I .. . . ., . . .. , otlrltvvlifit .l , . . . I”..£ .8: . . I {11¢ I , , . . . {3.15. {.1 . .fi1I.I . .. . In? 13. I . a . . 1w 3.... .. . _..w..a§w........ .s 5.9:: u . . I. 1.9..indhuflzn 2 wwwowuuk. ‘i..flur.LJ.uuvm\lug 33.17331 11,”! , . a «I! J- 4% . imdnhwxa. a... flank? . I . I .. I. I all-Z. ” I g by {Ari III IV . o .. . .5... kn.fl.uV.a.\s1! vn . .Ib.) .1 . {£00. . .. . . ahvnxuv 4-r...a:fl..da!. 1v}... .4.‘ .I .. p. A . 5w»: knuuflbusp ..I.......1..,.b.«t..v.....ks..r I I \Q’ A.| ll $32.4. uvvJquhw. :. ... db.bvv.‘....c. 2.»..1 HI 1.... ::..AH¥......»..;.1......§- I a... .. 2 2 2. I I...» «a .123: . =2 . . a. . 2.1.1... -.. I.. 1.... . K .w It: I .V -l i . 4 . . . 1..Jt;H\$Vl...fl L1 . . . u . . .1: . I . . . , {th Hwy . , {I . . I. . . an..." I. . 1 . . . I .. , . a! , 3% «III; . utqlrwlfinmi. «tax... . z . . Mafiauflutu. .: ..A I ‘4' I" II ll‘VXVI V I! Kgfih‘. ‘Uu‘15h' o .. I -. >I ou.’vl‘ (as? £31.33... .gflvflhfiefiiInféznnfifi . I 33.1.. I {LaunLKaTIiv :2..I.rv.b.x}..5.u.dv.%¢wa .3 ., . I: an; . I. . I .v. .I, I1 3 I $\IIOJ"1« I .. . $ 1 Q. s... I , Iinv H. .5. n . . 34...}...9 1. 3.43.... t . . . ,. . a . I. I I :5! I: n. n I . s. ., 9 r If. {a it 1-4 2:2 Ifiw§:g . . . :3 . . qfidt. . in, . IL .2 .» . .. .w . . . .. 21..ng «Mannahlilw .v ...I FIIHJénPQI) 2 $2.- . . 1 -8: . z . Jwfifln x d»: I... . I “9th “VIII IHIIIMII‘I. I I“? IIIIIIII U. .. n (fit. vii Rani. I. .Icflnf o . fit .Ilo‘uplnl‘ I I ., .Inf.. .I . hulvv~ .. HI . v . ,. . .. : , . I u I . . . 2 . . u > . L I. . t . I I i“:1' - . n a I... .‘Jl’bl | MIC. . . I ’ I lllv'l \ID I. Ck! a‘l. - t 4 . lvl!|Il-I‘-.‘I1'AIH\ I-II- IIIII w? :31 many} III II . I 3, u“ .I 92‘ |’ 3?]: I}! . II‘ .31 . l}.- ‘ I ‘ o . . I... I.‘ I I l‘h‘ v . .I «.uIIIPK .IIOII I} JIIL IIIII .. . IIIII 31H. (uvkoo) II vIII2 . I I I V .LLVI I. .II I..I . fall I. unify-61.53.21 , . ii...) .uU-IFIIOI. A. u!“ «JV-fifldvtfli I . III—1:: It IR... III! I... v . I.V....I.Iol.ll»v “100.... ..u.. I . . I > a . . I. . a: u. I I I v 1 I. .. v . . . I v n u I n-Illtl‘ll ‘ . v. I a. I! 5%. v n» I- "‘ "L.! I: . ~00, VII-I.c . I‘IOII‘X . riv,ofl' . . Isiin-‘Oyv! v.l‘1tt‘:is. 3‘ In. . III‘\Irv-‘9JNN . \I .. , htts‘k.“(..f..- I . ~7...o‘0|)|asst II.~.II. - JI.|..\.I§..\II\~I . ‘8‘IIIVuu-V ‘va.wv . ‘1. . I I I I III LD'IVIII'JJHJWIIII‘AHHI n . I L, . _ ..I_..I ,I . Ltr I. :. I.‘I fl .1 ,O-.II I. l Illllls I- 3 1293 01570 7908 ii" 1Illillllllllllllllllllllllll'llllllilllllillI r This is to certify that the dissertation entitled "Structure and Magnetism in Cobalt/X, Fe/Si, and Fe/[FeSi] Multflayers" presented by Michael Ray Franklin has been accepted towards fulfillment of the requirements for Ph. D. degree in PhYSiCS (MM/7% («4/ Ma jorr professor Date July 22. 1997 M5 U is an Affirmative Action/Equal Opportunity Institution 0-12771 STRUCTURE AND MAGNETISM IN Co/X, Fe/Si, AND F e/ {FeSi} MULTILAYERS by Michael Ray Franklin A DISSERTATION submitted to Michigan State University in partial fiilfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy 1997 ABSTRACT STRUCTURE AND MAGNETISM IN Co/X, Fe/Si, AND Fe/{FeSi} MULTILAYERS by Michael R. Franklin Previous studies have shown that magnetic behavior in multilayers formed by repeating a bilayer unit comprised of a ferromagnetic layer and a non-magnetic spacer layer can be affected by small structural differences. For example, a macroscopic property such as giant magnetoresistance (GMR) is believed to depend significantly upon interfacial roughness. In this study, several complimentary structural probes were used to carefully characterize the structure of several sputtered multilayer systems-Co/Ag, Co/Cu, Co/Mo, Fe/Si, and Fe/{FeSi}. X-ray diffraction (XRD) studies were used to examine the long-range structural order of the multilayers perpendicular to the plane of the layers. Transmission electron diffraction (TED) studies were used to probe the long-range order parallel to the layer plane. X-ray Absorption Fine Structure (XAF S) studies were used to determine the average local structural environment of the ferromagnetic atoms. For the Co/X systems, a simple correlation between crystal structure and saturation Michael Ray Franklin magnetization is discovered for the Co/Mo system. For the F e/X systems, direct evidence of an Fe-silicide is found for the {FeSi} spacer layer but not for the Si spacer layer. Additionally, differences were observed in the magnetic behavior between the Fe in the nominally pure Fe layer and the Fe contained in the {FeSi} spacer layers. in memory of Dr. Carl L. F oiles iv ACKNOWLEDGEMENTS I wish to thank Dr. Carl F oiles for his invaluable guidance and support during this project. He was more than just a research advisor; he was also a good friend. I also wish to thank Dr. Jerry Cowen for several enlightening conversations about magnetism and Dr. Peter Schroeder for his final review of this dissertation. Thank you to all the faculty, staff, and graduate students in the Physics Department at MSU. So many have helped me more than they know. Special thanks are given to Reza Loloee and Vivion Shull for their many hours spent helping me with this project. Special thanks to Paul Holody for our many discussions of physics. Also, special thanks to Dr. Jules Kovacs and especially Stephanie Holland for helping me with those administrative headaches. Also, ll must thank my parents for the many years of selfless support and wise instruction they have given me. I could never have reached this point without them. Many thanks are also given to my wife Kathy for her sacrifices of time as well as for her never-ending support. Most of all, I wish to thank God for giving me such a wonderful opportunity. I pledge to use this blessing to serve my fellow man. Finally, I wish to thank the Department of Physics and Astronomy and the Center for Fundamental Materials Research for their financial support during my time at MSU. TABLE OF CONTENTS 1. INTRODUCTION ................................................................................................. 1 2. SAMPLE FABRICATION ................................................................................... 3 2.1 Sputtering Process ......................................................................................... 3 2.2 Sputtering System Description ..................................................................... 5 2.3 System Preparation and Cleanliness ............................................................ 7 2.4 Typical Operating Procedures and Conditions .......................................... 7 3. X-RAY DIFFRACTION ..................................................................................... 10 3.1 Introduction ................................................................................................. 10 3.2 Bragg's Law ................................................................................................. 11 3.3 Powder Diffractometry ............................................................................... 13 3.3.1 Diffraction Patterns ............................................................................. 14 3.3.2 The Rigaku Rotating-Anode Powder Diffractometer ......................... 19 3.4 Typical Measurements ................................................................................ 24 3.5 Analysis ......................................................................................................... 24 3.5.1 Determination of Planar Spacings ...................................................... 24 3.5.2 Determination of Bilayer Spacing ...................................................... 25 3.5.3 Determination of Coherence Length (Scherrer's Equation) ................ 27 vi 4. TRANSMISSION ELECTRON DIFFRACTION ............................................ 31 4.1 Introduction ................................................................................................. 31 4.2 Bragg's Law Modification .......................................................................... 33 4.3 Diffraction Patterns ..................................................................................... 34 4.3.1 The Ewald Sphere ............................................................................... 34 4.3.2 Single Crystal and Polycrystalline Diffraction Patterns ..................... 35 4.3.3 Textured Diffraction Patterns ............................................................. 35 4.4 Transmission Electron Microscopy ........................................................... 38 4.4.1 Conventional TEM ............................................................................. 38 4.4.2 Image and Diffraction Pattern Formation ........................................... 41 4.4.3 VG HBSOl FESTEM .......................................................................... 46 4.5 Sample Preparation and Typical Settings of the FESTEM ..................... 49 4.6 Analysis ......................................................................................................... 50 4.6.1 Structural Determination .................................................................... 50 4.6.2 Compound Identification .................................................................... 51 5. X-RAY ABSORPTION FINE STRUCTURE ................................................... 54 5.1 Theory ........................................................................................................... 54 5.1.1 Physical Origin ................................................................................... 55 5.1.2 The XAF S Equation ........................................................................... 55 5.2 Experimental Procedures ............................................................................ 57 5.2.1 The Synchrotron ................................................................................. 58 5.2.2 XAF S Measurement Techniques ........................................................ 59 vii 5.3 MacXAFS Analysis ...................................................................................... 62 5.3.1 Data Conversion (Raw -—) XMU Files) .............................................. 65 5.3.2 Normalization and Background Removal (XMU —> CH1 Files) ........ 65 5.3.3 Fourier Transform (CHI —) RSP Files) .............................................. 68 5.3.4 Inverse Fourier Transform (RSP —> ENV Files) ................................ 69 5.3.5 Least-squares Fit ................................................................................. 70 6. CO/X MULTILAYERS ...................................................................................... 72 6.1 Introduction ................................................................................................. 72 6.2 Co/Ag ...... - ...... - -- - ...... - -- 73 6.3 Co/Cu ............................................................................................................ 74 6.3.1 Sample Fabrication ............................................................................. 76 6.3.2 XAF S Results ..................................................................................... 76 6.3.3 XRD and AXRD Results .................................................................... 80 6.4 Co/Mo ........................................................................................................... 86 6.4.1 Samples Fabricated ............................................................................. 88 6.4.2 Magnetic Studies ................................................................................ 88 6.4.3 XRD Results ....................................................................................... 91 6.4.4 TED Results ........................................................................................ 94 6.4.5 XAF S Results ..................................................................................... 94 7. FE/SI AND FE/{FESI} MULTILAYERS ....................................................... 104 7.1 Introduction ............................................................................................... 104 viii 7.2 Samples Fabricated ................................................................................... 107 7.3 Magnetic Studies ........................................................................................ 107 7.4 XRD Results ............................................................................................... 110 7.4.1 Low-angle XRD ................................................................................ 110 7.4.2 Higher-angle XRD ............................................................................ 112 7.5 TED Results ............................................................................................... 115 7.6 XAFS Results ............................................................................................. 119 8. SUMMARY AND CONCLUSIONS ................................................................ 122 8.1 Co/Cu Multilayers ..................................................................................... 122 8.2 Co/Mo Multilayers ..................................................................................... 122 8.3 Fe/Si and Fe/{FeSi} Multilayers ............................................................... 123 8.4 Future Investigations ................................................................................. 124 8.4.1 Co/X Multilayers .............................................................................. 124 8.4.2 Fe/Si and Fe/{FeSi} Multilayers ...................................................... 125 ix Table 2-1: Table 6-1: Table 6-2: Table 6-3: Table 6-4: Table 7-1: Table 7-2: LIST OF TABLES Deposition rates for various targets. .................................................................. 8 Saturation magnetization and fitted EXAF S results.5 N values are estimated to be accurate to within i 15%. GMR values are current-in-plane data at 5K from PA. Schroeder—the percent change is relative to the resistance at or near magnetic saturation. ................................................................................ 74 EXAFS results for Co/Cu samples. The upper portion of the table contains results for the sputtered samples of the present study. The lower portion of the table summarizes the results reported by Pizzini, et al. for evaporated samples.15 Since Pizzini, et a1. fixed the coordination number N at 12, the data analysis for the present study treated N both as a fixed parameter and as a variable to be fit. .......................................................................................... 78 Summary of d-spacings from XRD and AXRD data. I<> denotes the dominant line; 1' denotes the satellite located just below the dominant line while F denotes the satellite just above it. A is the measured value of the bilayer spacing. The * indicates unusually large intensities and the ** indicates the sample with a 60A Fe buffer and a 50A Cu cap. (The buffer is the first layer sputtered; the cap is the last). The procedure for calculating positions is described in the text. ....................................................................................... 82 XRD structural results for Co/Mo multilayers. denotes a single line interpreted as a composite line composed of the Mo<110> line and 3 Co line. The relative error for these d-spacings is $0.003 A. The * denotes a broad, weak line which is more likely an indication of the nearest neighbor distances in an amorphous sample. ................................................................................ 93 Magnetic properties of Fe/{FeSi} multilayers as a function of spacer layer thickness. The M5,. data were obtained with the DC SQUID, and the Hsat and Mr/Msat data, with the AC SQUID. ............................................................... 108 TED Bragg lines observed in Fe/Si samples. The lines are identified by their Miller indices. Y indicates that the line is observed, and N that it is not. ...1 17 Figure 2-1: Figure 2—2: Figure 3-1: Figure 3-2: Figure 3-3: Figure 3-4: Figure 3-5: Figure 3-6: Figure 3-7: Figure 3-8: Figure 4-1: Figure 4-2: Figure 4-3: Figure 4-4: Figure 4-5: Figure 4-6: Figure 4-7: Figure 4-8: LIST OF FIGURES Schematic diagram of the sputtering process. .................................................. 4 Sputtering system schematic (courtesy of P. Holody) ...................................... 4 X-rays diffracted from atomic planes. ............................................................ 12 Diffractometer layout. .................................................................................... 14 XRD diffraction pattern for Co/Mo (14.4A/1 1.2A). ...................................... 15 Harmonics for Co/Mo (14.4A/11.2A) shown in Figure 3-3. ......................... 17 The average Bragg line and its satellites for Co/Mo (14.4A/1 1.2A). ............ 17 Schematic diagram of the Rigaku rotating-anode powder diffractometer. ....20 Typical A1 sample holder used in XRD. ........................................................ 22 Focusing geometry for (a) small incident angle and (b) large incident angle.24 Transmission geometry in TED (after von Heimendahl). .............................. 32 The Ewald sphere construction in a) XRD and in b) TED ............................. 34 Portion of the reciprocal lattice for a (001) fibre-textured bcc film (after Tang and Thomas). .................................................................................................. 36 STEM imaging mode (after Watt). ................................................................ 39 Schematic of the primary components of a typical conventional TEM. ........ 40 The objective aperture blocks the diffracted rays thus providing contrast in the TEM image. .................................................................................................... 42 Schematic of (a) BF - and (b) DF-imaging. Io = primary beam, It = transmitted beam, Is = weakly scattered beam, and Id = diffracted beam. ......................... 42 Image formation in a conventional TEM. The rays for the diffracted beams are blocked beyond the plane of the objective aperture but are included here for illustration (after Crimp). .......................................................................... 44 Figure 4-9: Diffraction pattern formation in a conventional TEM (after Crimp). ............ 45 Figure 4-10: Schematic diagram of the primary components used in TED for the VG I-IBSOI FESTEM. ........................................................................................... 47 Figure 5-1: Schematic diagram of a Lytle detector. .......................................................... 61 xi Figure 5-2: Figure 5-3: Figure 5-4: Figure 6-1: Figure 6-2: Figure 6-3: Figure 6-4: Figure 6-5: Figure 6-6: Figure 6—7: Figure 6-8: Figure 6-9: Raw data from a fluorescence measurement of a 2000A sputtered Co film. .63 CHI data for the same 2000A sputtered Co film shown in Figure 5-2. ......... 63 RDF for the same 2000A sputtered Co film shown in Figure 5-2. ................ 64 Normalized saturation magnetization vs. nominal Co layer thickness for Co/Mo multilayers with tMo = tCo- In addition to data from the present study, the data of Wang, et a1.20 and Sato19 are presented. ........................................ 89 Field dependence of the normalized magnetization for the a) 14ML/14ML and b) 2ML/8ML Co/Mo multilayers on A1203 substrates. The diamagnetic contribution of the substrate is fit with a straight line to determine the y- intercept thus providing the saturation magnetization of the multilayers. ...... 90 Comparison of the higher-angle XRD data for the 5ML/5ML, 10ML/10ML, and 28ML/14ML Co/Mo multilayers. ............................................................ 92 TED line scan for Co/Mo (5 ML/SML). ........................................................ 95 TEM image of Co/Mo (7 ML/SML). The nominal magnification is 50 K. ..95 TED line scans from different regions of the same Co/Mo (7ML/5ML) sample illustrating the varying degree of crystallinity. ................................... 96 a) TED line scan for Co/Mo (7ML/7ML) and b) TED line scan for Co/Mo (IOML/IOML) ................................................................................................. 97 Combined raw data from 5 fluorescence scans of the 7ML/7ML multilayer.98 Comparison of the CHI and RSP data for the 5ML/5ML and 7ML/7ML samples clearly shows the transition from an amorphous to a crystalline structure. ......................................................................................................... 99 Figure 6-10: Saturation magnetization and number of nearest neighbors as a function of Figure 6-1 1 Figure 7-1: nominal Co layer thickness. The solid data points are for samples of equal layer thickness. The unfilled data points are for the 28ML/14ML and 2ML/8ML multilayers. ................................................................................. 100 : Comparison of the CH] and RSP data for the Co standard and all of the Co/Mo multilayers included in the XAF S study ........................................... 101 Field dependence of the magnetization normalized to the volume of the nominally pure Fe layers for the a) {Fe2ASi2A} X 4 and b) {Fe2ASi2A} X 7 spacer layer samples on A1203 substrates. These AC SQUID data give slightly larger saturation magnetizations than the DC SQUID data (see Table 7-1). ............................................................................................................... 109 Figure 7-2: Measured bilayer spacing as a function of nominal Si layer thickness for F e/Si multilayers. The solid line indicates nominal bilayer spacings. .................. 111 xii Figure 7-3: Figure 7-4: Figure 7-5: Figure 7-6: Figure 7-7: Coherence length vs. nominal Si layer thickness for both Fe/Si and F e/ {F eSi} multilayers on A1203 substrates. The horizontal line indicates a constant value .............................................................................................................. 1 1 1 a) Planar spacing as a function of spacer layer thickness for the F e/Si samples and b) a comparison of the higher-angle XRD data for the tgi = 10A and ts, = 20A multilayers ............................................................................................. 113 a) Planar spacing as a function of spacer layer thickness for the Fe/{FeSi} samples and b) a comparison of the higher-angle XRD data for the {FeZASiZA} X 3, X 6, and X 10 multilayers. .............................................. 114 TED line scans for Fe/Si samples with nominal Si layer thicknesses of a) 12A and b) 30A. ................................................................................................... 116 TED line scan for the {Fe2ASi2A} X 10 spacer layer sample with the two non-bcc lines indicated by the arrows. .......................................................... 118 xiii CHAPTER 1 INTRODUCTION Magnetic coupling in artificially-layered metallic systems known as multilayers is a current topic of debate. Since many of these systems have potentially important technological applications, the nature of the coupling mechanism needs to be thoroughly understood. Specifically, certain multilayers formed from the repeating of a bilayer unit comprised of a magnetic layer and a non-magnetic spacer layer show oscillations between ferromagnetic and antiferromagnetic coupling which accompanies giant magnetoresistance (GMR) effects. One hypothesis for the coupling mechanism is an RKKY-type coupling which is mediated through the spacer layer. The period of such a mechanism should be dependent on the Fermi surface of the spacer material; however, current experimental results show that the observed period is very insensitive to the specific metal used as a spacer. This result along with some other observations has led many investigators to claim that small structural changes may greatly affect the coupling mechanism. As a result of such claims, a very careful structural characterization of several systems including Co/Ag, Co/Cu, and Co/Mo has been performed. These different metallic spacer layers provide a variety of crystal structures. In addition, Fe/Si which has a non-metallic spacer has been studied for comparison to the other systems. The study of Fe/{FeSi} was included in an attempt to understand the Fe/ Si system better. Several types of structural characterizations have been done. Standard x-ray diffraction (XRD) has been used for a direct probe of the structure perpendicular to the plane of the layers, and transmission electron diffraction (TED) has been used as a complementary in-plane probe. Both of these methods have provided average information about each multilayer sample as a whole; however, these probes alone fail to provide a complete picture of the interfacial structure. For additional information, X—ray Absorption Fine Structure (XAF S) measurements have been used to determine the average local environment about the magnetic atoms to provide information about the interfacial regions. Correlations between the coordination number as measured relative to a standard and saturation magnetizations have been observed which support the idea that the detailed structure does indeed affect the magnetic behavior of these systems. Any conclusions drawn about the actual structure of the multilayers are model dependent; therefore, several different probes which provide complimentary information about the structure of the multilayers were chosen. Only in this way could the number of possible models be significantly reduced since a satisfactory model must be consistent with the results of every probe used. Since the reasons for researching the above systems differ, they are discussed individually in each introductory section prior to the experimental findings. Fairly detailed descriptions of each structural probe are also presented at a level hopefully requiring little knowledge beyond that obtained at the undergraduate physics level. Finally, the results of the experiments are given along with the conclusions drawn. CHAPTER 2 SAMPLE FABRICATION All sputtered samples were made at Michigan State University using a dc-magnetron sputtering system and were subsequently stored in a vacuum dessicator at approximately 0.1 torr to limit surface oxidation. Several different substrates were used including crystalline silicon and sapphire for XRD, magnetic, and synchrotron studies as well as sodium or potassium chloride substrates for TED studies. No definitive evidence was found that the choice of substrate had any significant effect on the multilayer structure. In the remainder of this chapter, a detailed description of the sputtering system and its operation will be given. Other sources for descriptions of the sputtering system which include some excellent detailed drawings are the dissertations by P. Holodyl and ML. Wilson.2 2.1 Sputtering Process Sputtering is a method of deposition in which ionized gas atoms collide with the material to be deposited—the target or source—causing the ejection of individual atoms which then accumulate on a substrate. This process is shown schematically in Figure 2-1. During sputtering, electrons are thermally activated and then accelerated to ionize the gas (usually Ar) which is sufficiently confined to create a plasma. The gas ions are then accelerated toward the target by holding it at a large negative potential. Upon collision with the source material, the ions transfer their kinetic energy to individual target atoms causing them to be ejected in all directions above the plane of the source because the incoming ions have undergone multiple collisions and consequently come in from all 3 a 6;) G? . <94 ‘ Figure 2-1: Schematic diagram of the sputtering process. stepping motor substrate holder atmosphere UI-IV shaft SPAMA plate \ Chimney _ Chimney with hole ' without hole - - . , sputtermg gun Figure 2-2: Sputtering system schematic (courtesy of P. Holody’). directions. A substrate is properly positioned to collect the target atoms. By controlling the time the substrate is positioned above a target and by monitoring the rate of accumulation on the substrate, the thickness of a given target material is specified. Moreover, by positioning the substrate over different targets for alternating set periods of time, artificially- layered metallic systems are created. 2.2 Sputtering System Description The dc-magnetron sputtering system shown in Figure 2-2 is housed in a stainless steel cylinder. At the bottom are four L. M. Simard "Tri-Mag" triode guns which as measured from the target center are placed 14 cm from the cylinder center and spaced 90° apart. Since all of the samples made consisted of a repeated bilayer, only two guns were required. Each gun assembly includes a filament and anode to provide the necessary accelerating electrons to ionize the sputtering gas. The gas used was always ultra-high purity Ar (99.999%) which is passed through a LNz cold trap, then a gas purifier, and finally forced through several small closely-spaced openings at the gun base. The target material is mounted on a Cu base which is subsequently attached to a Cu block containing circulating water to remove the tremendous heat buildup caused by the sputtering process. Both the base and the source are held at potentials of -400 to -600 V to accelerate the Ar+ ions. A grounded aluminum ring surrounds the target to limit the ejection of atoms from other areas. At the very top of the gun, a magnetic shield is positioned such that in combination with the two magnets at either end of the gun, a magnetic field is created to confine the Ar+ ions and create a plasma just above the target. The guns are covered by a manually—controlled rotating chimney assembly wrapped in aluminum foil for ease of cleaning. Each gun requires two chimneys. The first is completely covered with foil to help prevent contamination when the shutter (described in the following paragraph) which is normally protecting the sample must be opened for a brief period prior to sample fabrication. The second is rotated into position once the sample is actually ready to be made. It has a small opening approximately 5 cm in diameter to limit system contamination. Above the chimney assembly lies the computer-controlled SPAMA (Substrate Positioning And Masking Apparatus) plate. Eight sample holders are mounted on it—each of which contains two sample substrates. The sample holders consist of two openings which are flared out at the bottom to be smaller than the substrates. Also, a small Cu block is placed above the substrates under a slight amount of pressure and connected to the SPAMA plate for temperature control. The SPAMA plate itself is cooled using high- pressure N2 gas cooled by LNz. Underneath each holder are manually-rotated shutters which help prevent contamination of the samples either before or after manufacture. They may have either one or two openings. Those with two are used when the same sample is desired on two different substrates so that they may be made simultaneously. In addition to the sample holders, two quartz film thickness monitors have openings on the SPAMA plate. These measure the resonant frequency of the quartz disk which decreases as mass is added to it. Given information about the density and acoustic impedance of the target, this change in the resonant fiequency can then be used to calculate the film thickness accumulated over a given time interval, i.e., the deposition rate. It is measured for each target prior to, periodically during, and after each run so that accurate values may be entered into a computer which is programmed to rotate the SPAMA plate back and forth between targets holding the substrates in place over each target for the appropriate time interval. Configurations other than that depicted in the preceding are also possible. These are described in detail elsewhere.4 2.3 System Preparation and Cleanliness First, all of the parts mounted on the SPAMA plate were removed. Then the aluminum gun casings, target rings, and those parts mounted underneath the plate and consequently exposed to the target beam (the parts on the top of the SPAMA plate were not cleaned) were placed in a bath consisting of 3 parts HNO3 acid (70% soln.) /1 part H20, rinsed in water, and scrubbed with a wire brush. The SPAMA plate was also chemically etched with the acid solution afterward but not immersed in it. When this procedure was not sufficient for cleaning, it was repeated or a small amount of HF acid added for those parts made of stainless steel. After rinsing with water, all of these parts except the SPAMA plate were cleaned ultrasonically first using acetone and then ethyl alcohol. The plate itself was too large for this procedure and consequently was only rinsed with these chemicals. Additionally, the Si and A1203 substrates were treated using this ultrasonic cleaning method while the salt substrates were cleaved just before insertion into the sample holders. Assembly of the system was performed while wearing cotton and/or latex gloves. 2.4 Typical Operating Procedures and Conditions The vacuum chamber was initially pumped down using an Edwards E2M-18 mechanical pump with a molecular sieve trap to limit oil backstreaming and further evacuated using a CTI Cryo-Torr 8 cryopump. Additionally, the system was typically baked overnight for approximately 16 hours at 60°C enabling base pressures on the order of 10'8 torr to be achieved as measured using a Dycor M100 Quadrupole residual gas analyzer. During a run, the Ar required for plasma formation was admitted to the system at a pressure of 2.5 mtorr. The substrate temperatures were monitored using two thermocouples in thermal contact with two of the substrates and were typically maintained between -5 0°C and -20°C. Only one shutter was open at a given time, and the covered chimneys remained closed over all of the targets until just prior to deposition. They were then opened during deposition and closed again immediately afterward. Deposition rates were generally chosen to be the maximum values possible without losing the Ar plasma or shorting out the target. These values varied depending upon the source material. Table 2-1 shows the range of deposition rates associated with each target. The typical time to fabricate one sample was 15 minutes with the exception of the F e/ {F eSi} samples so run times were kept to a minimum—usually about 2 hours. In the F e/ {F eSi} case, sample times were approximately 30 minutes each, and these runs lasted 3-4 hours. Table 2-1: Deposition rates for various targets. Deposition Rate Target (A/s) Co 3.4-9.0 Cu 3.4-14 Fe 4.0-6.6 Mo 7.9 Si 2.4-4.3 ‘P. Holody, Ph.D. dissertation, Michigan State University (1996). 2ML. Wilson, Ph.D. dissertation, Michigan State University (1994). 3P. Holody, ibid. 4J.M. Slaughter, W.P. Pratt, Jr., and RA. Schroeder, Rev. Sci. Instrum. 60 (1), 127 (1989). CHAPTER 3 X-RAY DIFFRACTION 3.1 Introduction X-ray diffraction (XRD) results from the scattering of x-rays by electrons in atoms. When atoms are arranged periodically on a lattice, definite phase relationships exist between the scattered rays. This scattering occurs in all directions, but in most directions destructive interference occurs between the scattered rays coming from different atomic sites. Only in a select few directions does constructive interference occur—this reinforcement of scattered rays is known as diffraction. The solid state physics texts by 2 contain brief introductory descriptions of Myers1 and also by Ashcroft and Mermin crystallography and XRD while Cullity’s classic text3 is a more in-depth reference. The manner in which x-rays are diffracted is unique to a given crystal structure. So analysis of a resulting diffraction pattern whether recorded on photographic film or electronically with detectors will reveal structural information for a particular sample. If a sample consists of a single crystal or a random distribution of one type of crystal structure, then the structure of that sample can be uniquely determined. Complications may arise for systems consisting of more than one crystal structure—multilayers are often such systems. There are many different methods for observing how x-rays are diffracted from a particular sample, and the resulting diffraction pattern is determined not only by the crystal structure of the sample but also by how the diffraction pattern is recorded. The method chosen depends upon the characteristics of the sample and upon what information 10 11 is to be determined. For example, the back-reflection Laue method can be used to determine the orientation of a single crystal, the rotating-crystal method can be used to determine the unknown structure of a single crystal, and the Debye-Scherrer or powder method can be used to determine the unknown structure of a polycrystalline sample or a powder of single crystals. All of these techniques record the diffracted x-rays on photographic film with single crystal patterns appearing as spots and polycrystalline patterns appearing as rings or portions of rings depending upon the film size. The use of film allows quick measurements but limits the accuracy of measured intensities. Other methods such as the use of a powder diffractometer utilize some type of counter to more accurately determine intensities. In the remainder of this chapter, a discussion of Bragg's Law, a description of powder diffractometry, and the analysis of typical multilayer diffraction patterns will be given. In this chapter, it is assumed that the reader has a basic understanding of crystallography including Miller indices and the reciprocal lattice. If not, the previously mentioned texts should be consulted. 3.2 Bragg's Law The atoms of any sample can be envisioned as lying on a set of atomic planes as in Figure 3-1. In this diagram, the incident x-rays strike the set of planes at an angle 6. Due to conservation of energy and momentum requirements, the diffracted x-rays must "reflect" from the sample at an angle 0 such that the scattering angle is 20. The terms "reflect" and "reflection" are often used in reference to x-rays diffracted from atomic planes even though no reflections actually take place. Using basic trigonometry, it is easy to determine the "Bragg condition" which is the 12 m 9 9 a v v i 9 9 d. A C B ll 29 Figure 3-1: X-rays diffracted from atomic planes. condition for diffraction to occur. The path difference for the two rays shown in Figure 3-1 is given by the length of the two line segments fl and If which equals 2d'sin0. The necessary requirement for constructive interference is that this path difference equals an integral number of wavelengths. The result is the well-known Bragg equation: 2d'sin9 = nl (Eqn. 3-1) where d' is the interplanar separation typically referred to as the "d-spacing", 9 is the angle between the incident x-ray and the sample, A is the wavelength of the radiation, and n is the order of reflection which can take on any integral value. Typically, this equation is written with d = d'/n such that all reflections can be considered as first order. In other words, an nth order reflection from the (hkl) planes of spacing d' can be regarded as a first order reflection from the (nh nk n1) planes of spacing d, either real or fictitious, such that the result is the more familiar form of the Bragg equation: 13 2d sine = A (Eqn. 3-2) Though Bragg's Law specifies the conditions for diffraction to occur, it does not provide any information regarding the amplitude of the diffracted beams from a given set of planes. The amplitude depends upon the crystal structure, so that certain families of planes have zero amplitude diffracted beams referred to as forbidden lines. (The reader should consult one of the texts mentioned in the introduction about the structure factor if unfamiliar with this result). The non-zero ones are referred to as the allowed lines or peaks for that crystal structure and are unique to it; therefore, a sample's structure can be determined by identifying the allowed lines. For example, the first 5 allowed lines of a body-centered cubic (bcc) structure, (110), (200), (211), (220), and (310), and the first 5 allowed lines of a face-centered cubic (fcc) structure, (111), (200), (220), (311), and (222) provide a distinction between the two structures. Also, a useful resource in identifying structures from their observed lines is the Powder Diffraction File.4 3.3 Powder Diffractometry A powder diffractometer consists essentially of an x-ray source, a sample mount, a detector, and some slits. For the standard 0-29 reflection geometry shown in Figure 3-2, the x-ray source remains fixed while the sample and the detector are rotated synchronously during a scan such that the detector moves through twice the angle that the sample moves through. The incoming x-rays are incident at an angle 9 with the sample plane and the detector is located at 29; consequently, diflraction occurs for those planes which are nearly coincident with the sample plane. 14 detector Figure 3-2: Diffractometer layout. 3.3.1 Diffraction Patterns A diffraction pattern recorded by a powder diffractometer is typically displayed as the number of counts/second (N) vs. the scattering angle (29). It varies fi'om large regions in 29 where N is very small and fairly constant due to background radiation to much smaller regions where peaks (commonly called Bragg lines or peaks) are observed in N which result from diffraction by the sample. In the remainder of this section, the diffraction patterns associated with a single crystal, a polycrystalline sample, and a multilayer are discussed. If a single crystal were placed in a powder diffractometer, most likely nothing would be seen in the diffraction pattern since the likelihood that any set of planes would be properly aligned to satisfy the Bragg condition would be small. In a polycrystalline or a 15 2 _ I I r I I I f T I f I r I r I I I d 103 E ave. Bragg line J A * I U) “ .1 3:: 3 “ i t: 2 r . . 1 =5 102 i harmonics satellrtes _? g : : V 3 _ 2nd order effects 3? 2 r 1 'c'Z T : I: 101 g .2 0 : I E * ll . ‘ fl 3 a l ‘ ‘ _, 2 r u ' l ' ‘ i t ‘ , j ”’ I : 100 r ‘ ' ' g l L L I l l A J n M 1 1 I 1 4 1 i I 50 6O 70 80 90 100110 29 (deg) O 10 20 3O 40 Figure 3-3: XRD diffraction pattern for Co/Mo (14.4A/11.2A). powdered single crystal sample in which all possible planar orientations are contained in one sample, all allowed Bragg lines corresponding to the crystal structure will be observed up to the limit of 29 allowed by the diffractometer which is typically around 135°. If the crystals contained in the sample were infinitely large and the diffractometer ideal, these lines would be spikes with no width. Departures from these conditions cause these spikes to be broadened. The explanation for this broadening is given in section 3.5.3. However, an XRD pattern for a typical sputtered multilayer such as that shown in Figure 3-3 does not contain all of the allowed Bragg lines. Instead, it has only one Bragg line (sometimes also a weaker second order line—for example, both the (110) and the (220) lines might be observed for a bcc structure) which can be identified as belonging to the two constituent components making up the bilayer unit if the layers are thin enough l6 and if the components have comparable d-spacings for their densest set of planes. These planes are usually the planes with the greatest density of atoms corresponding to the first allowed line for cubic systems. In such a case, the Bragg line will typically be a weighted average of the two d-spacings which depends on the relative atom concentrations; otherwise, 2 Bragg lines may be observed—one for each component. This result is consistent with the presence of fibre texture. When a sample possesses fibre texture, then only one family of planes (a set of symmetrically equivalent planes) is oriented parallel to the substrate; consequently, only lines associated with those planes can be observed in the standard 9-29 geometry. Of course, these planes are not perfectly parallel to the sample plane, but if the vast majority of planes which lie nearly parallel to the sample plane belong to this family, then the sample is said to possess strong fibre texture. This result is expected since it is well known that sputtered metallic atoms typically align themselves such that only the densest set of planes lie parallel to the substrate. In addition to the appearance of the above-mentioned lines corresponding to the d- spacing of the individual components, additional lines are found in the diffraction pattern. These lines result from the repeating bilayer unit of the multilayers and manifest themselves in two different ways. First, at low angles, lines are observed which are referred to as harmonics (see Figure 3-4). These lines obey Bragg's Law which is usually rewritten in this slightly modified form: 2Asin9 = ml (Eqn. 3-3) l7 I I * I I n I F r T ‘ 4 ._ q 3~ . 27 1 A » . i3 ' . a 102 E harmonics E :3 E 3 ‘5 4t * r-t 3 ~ . 3 2 1 r: ’ ‘4 OF! 101 : 1 CD : : C I. . 2 ,~ - .5 3 — . 2 1 100 : j '- l I l L I I L l T O 2 4 6 8 10 12 14 29 (deg) Figure 3-4: Harmonics for Co/Mo (14.4A/11.2A) shown in Figure 3-3. 2 > I I I . I ' I ' T I I I I I I ' I j 103 E ave. Bragg line : A E satellite 3 t: 2 r i :3 t , . . 102 L satellrte : .D . : g ” satellrte I v 3 : 5‘ 2 r 1 '53 f ‘. t: 101 5 a O : 2 u _ - t: - - 1-1 3 a _ 2 r 1 100 - 34 36 38 40 42 44 46 48 50 52 29 (deg) Figure 3-5: The average Bragg line and its satellites for Co/Mo (14.4A/11.2A). 18 where A is the bilayer spacing (sometimes referred to as the compositional modulation wavelength), the m is now included so the observed lines will be denoted as lst order, 2nd order, etc., and the remaining quantities are the same as before. Typically, 2-6 of these peaks are seen before their intensities sufficiently decrease from varying average bilayer spacings, thickness fluctuations within layers, etc., to the point that they are no longer observable. Second, additional lines referred to as satellites are seen near the Bragg peaks as shown in Figure 3-5. These satellites can be associated with both the first and second order peaks. To fully understand why satellites occur only near the Bragg lines, it is necessary to examine the diffraction pattern for an idealized multilayer consisting of 2 alternating crystalline regions repeated for a large number of cycles described perfectly by a step model, i.e., no thickness fluctuations or interdiffusion. The intensity of such a multilayer is given by the following:5 2 sinnanvl 2(sinrtflnada Mo]2 2[sinitt’nbdb /do] sin‘rwda /d0 b sinnt’db /d0 (Eqn. 34) sinnt’nada /d0)(sinrt€nb MOJ} 2ff 2 + a ”05“ “i sinttrt1,/ti0 ksinntdb/d, where Z is defined by the momentum transferq = 2m / d , N is the number of times the bilayer unit is repeated (the number of cycles), 11a and 11b are the number of atomic planes of material a and b respectively, da and db are the planar spacings, n = na+ nb , ndo = nada+ nbdb , and fa and fb are the atomic scattering factors. The frequency of the first term listed in parentheses is obviously much larger than 19 the second term in the curly brackets. So this first term is a rapidly oscillating function, and the second term is an envelope function governing the intensity of the first term. For relatively small 113 and nb, this envelope function consists of a single peak centered around the value of the average lattice spacing. As the individual layers become thicker, however, it becomes a bimodal distribution peaked around (1, and db. (Notice that this equation explains the presence of the one or two Bragg peaks mentioned previously on p. 15 where fibre texture is first discussed). The intensities of these satellites will be symmetric about the Bragg peaks only if both d3: db and 11,: nb. 3.3.2 The Rigaku Rotating-Anode Powder Diffractometer In the following paragraphs, the components of the Rigaku rotating-anode system used in this study will be discussed in detail. Its overall layout including the various slits is shown in Figure 3-6. Finally, the reason why this basic geometrical setup is chosen will then be explained. Housed in a tube, the x-ray source consists of a metallic target—the anode—and a source of electrons which are accelerated toward the anode. X-rays are generated both by the deceleration of the electrons which generates x-rays with a continuous spectrum of energies as well as by the ejection of inner core electrons which creates x-rays of specific energies. This last process occurs during the atom's transition from an excited state back to its ground state when it emits photons with energies related to differences in the energy levels of the atom. This radiation is characteristic of the target material and is used for XRD because it has a much greater intensity than the continuous spectrum. The most intense of these characteristic emissions is the Kor emission which occurs ’ sample t ~ ~ 4?- L_.. I SOLII'CC Soller slits ’ E/ RS monochromator RSm ‘ detector Figure 3-6: Schematic diagram of the Rigaku rotating-anode powder diffractometer. 21 when an electron moves from the L level to the K level where K and L respectively denote the lowest two electron energy levels. The Kor emission is separated out from the total emission spectrum since the determination of d-spacings requires a monochromatic source of x-rays, but the Kort and the Kong radiation is not resolved, so the effective unresolved wavelength for a Cu target is 1.5418A.6 In order to increase the intensity of the characteristic x-rays, the tube current is increased which increases the number of electrons striking the target; however, since most of the electron energy is converted into heat, the maximum current is limited by the maximum amount of heat which can be dissipated. One method for dissipating this heat is to rotate the target which serves as the anode of the tube—this is known as a rotating anode. In this manner, greater intensities can be generated. The Rigaku system uses a Cu rotating-anode. The type of source used in the Rigaku system is known as a line source. It emits an x-ray beam in the shape of a narrow rectangle which is diverging both vertically (perpendicular to the diffractometer plane) and horizontally. Consequently, slits are used to collimate the beam, i.e., to make the x-rays parallel, which is required for proper focusing of the x-rays. The slits marked DS and SS are thin vertical divergence slits which collimate the beam horizontally and should be the same size. Typically, the smallest slits available (1/6°) were used for DS and SS to improve resolution although with some loss in intensity. Though a given Bragg peak position is unaffected by these slits, they do determine the inherent minimum width of the peaks. Since the peak widths are used to determine the coherence lengths of the samples, the slits were chosen to minimize these instrumental broadening effects. The Soller slit assemblies which consist 22 Front Back Side Figure 3—7: Typical Al sample holder used in XRD. of thin metal plates parallel to the plane of the diffractometer collimate the beam vertically. After leaving the source, the x-rays pass through a Soller slit assembly and the DS slit and emerge to strike the sample. The sample is held in place by an Al sample holder which is fastened by a clip against a flat metal plate. In this manner, the sample is secured in place at the center of the diffractometer circle which has its origin at the sample center and has a radius defined by the source-to—sample distance (this distance must also equal the sample-to-RS slit distance). The sample holders use either a rigid stainless steel clip or a spring-like Cu clip to hold each sample in place. The holder shown in Figure 3-7 uses a stainless steel clip. The clip presses the sample against four protruding "feet" which were machined to within a tolerance of i0.025 mm each and made so that their back surfaces just coincide with the front surface of the holder. Oftentimes, several Al lines from the holder were observed in the XRD scans. Since these lines very closely matched the expected d-spacings for Al, the x-rays are clearly striking the main body of the sample holder which lies in the same plane as the sample. If 23 the x-rays were being diffracted from the feet which lie significantly in front of the plane, the observed d-spacings would be smaller than expected. As long as the same sample holder was used for each sample in a run, then these lines can serve as an internal consistency check. Any difliacting x-rays are then reflected from the sample and pass through the SS slit, then a second Soller slit assembly, and another divergence slit marked RS for "receiving slit". The x-rays then strike a graphite monochromator which consists of a single crystal oriented such that the Cu Ka radiation satisfies the Bragg condition and is "reflected" into the monochromator receiving slit (RSm), i.e., the monochromator selects the desired wavelength along with its harmonics. After passing through RSm, the x-rays enter the detector. It consists of a scintillating NaI crystal which converts the x-rays into visible light before entering a photomultiplier tube. As shown in Figure 3-8, the source (S) and the focal point (F) both lie on the difiractometer circle with the flat sample located at the center (0). This particular geometry is known as the Bragg-Brentano geometry and is chosen to maximize the intensity of the detected diffracted beam. The dashed line in both figures indicates the focusing circle. As long as S, F, and 0 lie on this circle as shown, then from simple geometrical considerations,7 all x-rays leaving the source and striking the sample will be focused at point F. In order to be perfectly focused, the sample would need to have the same radius of curvature as that of the focusing circle; however, this requirement cannot be reasonably satisfied since the radius of the focusing circle varies with incident angle 9. As a result, some broadening of the diffracted beam does occur though this can be limited by a narrow incident beam. 24 focusing circle \ diffractometer circle (a) (b) Figure 3-8: Focusing geometry for (a) small incident angle and (b) large incident angle. 3.4 Typical Measurements For each sample, a relatively quick scan was performed over a very large range of 29—at least well beyond the 2nd order peak of the first allowed line. These scans typically covered 6°/min in 29 with a step size of 006°. After this initial investigation, slower scans were performed in the regions of interest—typically the low-angle harmonics, the average Bragg line along with its associated satellites at first and second order, and any undetermined lines. Depending on the strength of these lines, the scans typically covered O.25°-1.25°/min with a step size of 0.02°-0.03°. 3.5 Analysis 3.5.1 Determination of Planar Spacings The d-spacing of a given x-ray line was determined using software to place a cursor at the peak center. The value of 29, the d-spacing, and the intensity were printed out on the diffraction pattern corresponding to this position. This seemingly inaccurate 25 determination of the d-spacing actually introduced an exceedingly small error. To justify this statement, the peaks of four F e/ {FeSi} samples were fit using a Gaussian distribution along with an additional constant parameter for the undetermined background. Comparison between these values for the d-spacings and those determined by "eyeballing" the peak center showed a maximum Ad = 0.0006A. From these comparisons, it seems quite conservative to estimate a maximum error of Ad < 0.001A introduced by this method of measurement for well-defmed, symmetrically-shaped peaks. By comparing the measured d-spacings with those listed in the Powder Diffraction Files, the lines were identified. Another source of error is deviation of the sample position from the center of the diffractometer circle; however, no relative error is introduced as long as the same sample holder is used for samples in a given run. Since correlations between the structure and the exhibited magnetic properties are desired, only relative differences are important. Thus no attempts were made to calibrate the d-spacings using a standard, although the previously mentioned Al lines from the sample holders were used as a check for self- consistency. 3.5.2 Determination of Bilayer Spacing The bilayer spacing may be determined by fitting either the harmonics or the satellites positions as described in the remainder of this section. For calculation of the bilayer spacing using harmonics, the earlier statement of Bragg's Law is incorrect. At small angles, the index of refraction should be included. Not doing so can introduce errors in A as large as 3% though this percentage amounts to less than one monolayer in 26 the multilayers studied. The modified version of Bragg's Law8 is written as: 5 2A(1- sin2 9) sin9 = m (Eqn. 3-5) where 5 is defined by the index of refraction n=1-5. For x—rays, n<1 by a small amount which is typically 5z10'5 for most metals. The corresponding value of 9 for the mth order harmonic was determined using the method described in the previous section. The remaining symbols have the same meaning as in Eqn. 3-3. Unfortunately, a fit of data taken with the Rigaku system cannot determine both A and 8 as free parameters. At low angles, the displacement of the sample from the exact center of the diffractometer circle prevents an accurate determination of 5; consequently, 5 was calculated theoretically and a linear least-squares fit used to determine A. For an alloy, 8 = cfia + (1 — c)25b where a and b are the individual components and c is the alloy volume fraction of material a.9 This treatment seems justifiable since this expression should be reasonably close to the actual value assuming that the amount of compounds formed at the interfaces is relatively small compared to the remainder of the bilayer unit. The determination of the bilayer spacing using the satellite positions is accomplished by using the following equation which is derived by combining Equations 3-2 and 3-3: m = — (Eqn. 3-6) where m is the order of the satellite with planar spacing d. A linear least-squares fit was 27 used to determine A even though the order of each satellite was not determined since A depends solely on the slope of the described fit. In spite of the less complicated fitting procedure, the typically smaller number of satellites makes the determination of A using satellite positions less reliable than using harmonic positions. As a result, unless stated otherwise, all bilayer spacings listed herein were determined using harmonics. 3.5.3 Determination of Coherence Length (Scherrer's Equation) The structural coherence length is the average crystallite size in a particular direction. For a standard 9-29 scan, the average dimension measured is the one perpendicular to the plane of the layers. It can be obtained from the following relationship between the breadth of a line and the average crystallite size originally derived by Scherrer: _ Kl _ LcosG B (Eqn. 3-7) where 7t and 9 have the same definitions as before, [3 is some measure of the Bragg peak width (often, simply referred to as the linewidth), L is the crystallite size, and K is a constant approximately equal to unity and determined by assumptions about crystallite shapes and the precise definitions of L and B. In this treatment, B will be defined as the linewidth at FWHM in radians from the diffraction pattern on the 29 scale. For K taken to be unity, L is the volume-averaged crystallite dimension perpendicular to the reflecting planes. For a complete description of the treatment given by Stokes and Wilson eliminating the assumptions about crystallite shape and Gaussian symmetry made by 28 Scherrer, see the XRD text by Klug and Alexander—an encyclopedic reference for XRD by polycrystalline and amorphous materials.10 Line broadening can result from at least two other sources. First, an inherent minimum instrumental broadening occurs which is mainly a function of the width of the divergence slits DS and SS as well as the Bragg angle 9. For DS=SS=l/6°, the instrumental linewidth B varies from 005° for 29 = 10° to 0.1 1° for 29 = 120°.H Clearly, the measured linewidth must be larger than this instrumental linewidth to have any chance of determining a coherence length; therefore, the smallest divergence slits available were used since the intensity loss did not prevent observation of the lines. Second, non-uniform strain can cause the broadening of Bragg lines. While uniform strain leads to the displacement of a Bragg line from its non-strained position since the average d-spacing is different, non-uniform strain produces several closely spaced peaks which cannot be resolved thus creating one broadened peak. This effect along with others not discussed here such as the presence of faults must also be considered when determining coherence lengths. Thus, L is really an "effective" coherence length and not actually a crystallite dimension though this designation will not be used anywhere else in the text. However, the large changes observed in L for the Fe/Si system are attributed to an actual change in the coherence length since the other broadening effects would not be expected to change so dramatically. The coherence length L was calculated from Scherrer's equation using the measured value of B determined graphically directly from the diffraction pattern. Again, this might seem to introduce large errors, but they are actually quite small. For the same Fe/{FeSi} samples and fitting procedure mentioned in Section 3.5.1, B was calculated and compared 29 with the results obtained from a graphical analysis. It was found that AB was never larger than 0.05° which results in a 9% error in L. In summary, conventional 9-29 XRD diffraction patterns contain lines associated with the repeated bilayer unit and only one or two lines associated with the structure of the individual components of the multilayer which is indicative of fibre texture. As a result, it provides information concerning the long-range order of the multilayers perpendicular to the plane of the layers, but it is insufficient to determine the in-plane and interfacial structures of the multilayers. Alternative structural probes must therefore be used. IH.P. Myers, Introductory Solid State Physics, (Taylor & Francis, Philadelphia, 1991). 2Neil W. Ashcroft and N. David Mermin, Solid State Physics, (Saunders College, Philadelphia, 1976). 3B.D. Cullity, Elements of X-ray Diffraction, (Addison-Wesley, Reading, Massachusetts, 1956) 4Powder Diffraction File: Inorganic, formerly published by the American Society for Testing and Materials (ASTM) and presently published by the JCPDS (Joint Committee on Powder Diffraction Standards)—International Centre for Diffraction Data in Swarthmore, PA is a compilation of diffraction data from various sources consisting of many volumes. New ones are added as data for new materials are acquired. 5Leroy L. Chang and BC. Giessen, Synthetic Modulated Structures, (Academic Press, New York, 1985), p. 49. 6This value is the weighted mean of the Kort and Kong wavelengths with Kort being weighted twice that of Kong as is customarily done since the ratio of the intensities is 2:1. 713.13. Cullity, ibid., pp.156-7. 8Klug and Alexander, ibid., p.101 9P.F. Micelli, Appl. Phys. Lett. 48, 24 (1986). 3O 10Harold P. Klug and Leroy E. Alexander, X-ray Diffraction Procedures for Polycrystalline and Amorphous Materials, (John Wiley & Sons, New York, 1974), pp. 689-90. ”ML. Wilson, Ph.D. dissertation, Michigan State University (1994), p. 20. CHAPTER 4 TRANSMISSION ELECTRON DIFFRACTION 4.1 Introduction In many ways, transmission electron diffraction (TED) is similar to XRD. Electrons are scattered by atomic nuclei much the same as x-rays are scattered by electrons. This scattering of electrons leads to diffraction which obeys Bragg's Law and to the same extinction laws, i.e., the allowed lines for a given structure observed in TED are the same as those in XRD. Also, the diffraction patterns are quite similar with single crystal patterns consisting of spots and polycrystalline patterns consisting of rings, but there are important differences as well. Because the interaction between electrons and atomic nuclei is much stronger than those between x-rays and electrons, diffraction intensities in TED are 106-107 times stronger than those observed in XRD, penetration depths in TED are approximately 100 times lessl, and double diffraction can lead to the presence of forbidden lines in TED diffraction patterns except in the foe and bcc systems. (Double diffraction occurs in these systems as well, but a diffracted beam which is diffracted again can create only other allowed lines in these systems). A typical beam cross section used in TED is also smaller by a factor of 106-107 than a typical one used in XRD, so the affected sample volume is about 109 times smaller in TED.l But perhaps the most significant difference between TED and XRD results from the fact that electron wavelengths are approximately 100 times shorter than x-ray wavelengths. Consequently, the Bragg angle is limited to approximately 0-2°, and the reflecting lattice planes are very nearly parallel to the primary 31 32 + Primary beam if ‘ Wit 7 - Sample 1‘ ”a .. ,,,,,, ,, / Diffracted // % Transmitted beam \ beam 1K4 L Screen _J’_4 R -71— Figure 4-1: Transmission geometry in TED (after von Heimendahlz). beam as shown in Figure 4-1. Since diffraction is limited to these planes, then the structural information provided by TED is parallel to the sample plane making it a good complement to the perpendicular structural information gained from conventional XRD. In the remainder of this chapter, the modifications made to Bragg's Law, instrumentation, sample preparation, diffraction patterns, and subsequent analysis will be discussed. If the reader would like additional information about transmission electron microscopy in general or about TED in particular, several texts can provide that information. The text by Wyatt3 provides a good overview of the practical aspects of electron microscopy (EM) and an easily understood, yet thorough presentation of the concepts involved in transmission electron microscopy (TEM). Additional texts in increasing order of completeness and difficulty in understanding are those by von Heimendahl4, Thomas and Goringes, and Hirsch, et al.6 The latter is considered to be the classic text on TEM. 33 4.2 Bragg's Law Modification As mentioned in the introduction, the de Broglie wavelength of electrons is much smaller than the wavelength of x—rays generated by a Cu target. For example, electrons with an energy of 100 keV have a wavelength of 3.704 pm which is about two orders of magnitude smaller than Cu K.ll x-rays. It follows that 7t/2d <<1 since d-spacings are on the order of A. From Bragg's Law (2dsin9 = 2.), sin9 = M2d<<1; therefore, 9 is always quite small. From Figure 4-1 in which the transmission geometry is shown, it follows that tan29 = M (R and L are defined below). This result combined with the small angle approximation (sin9 z 9 and 0039 z 1) leads to the modified version of Bragg's Law used in TED: Rd = AL (Eqn. 4-1) where R is the distance between the diffracted spot and the transmitted (undeviated) beam as measured on a screen or photographic film, (1 is the planar spacing, it is the electron wavelength, and L is the camera length, i.e., the distance between the screen or film and the sample. If there are lenses between the sample and the screen, L is not an actual physical distance but the equivalent distance if the same size diffraction pattern resulted with no intervening lenses. The product AL is often referred to as the "diffraction" or "camera constant" and must be determined using a suitable standard if accurate measurements of the d-spacings are to be made. 34 Ak Figure 4-2: The Ewald sphere construction in a) XRD and in b) TED. 4.3 Diffraction Patterns The best method for understanding diffraction pattems is to use the Ewald sphere construction and the reciprocal lattice. In TED, this particular method is made simpler than in XRD because the shortness of the electron wavelengths make the sphere nearly planar on scales comparable to those of the reciprocal lattice. In this section, the relationship of the Ewald sphere to the diffraction patterns for single crystals, polycrystalline samples, and textured samples including multilayers will be discussed. 4.3.1 The Ewald Sphere In Figure 4-2, the construction of the Ewald sphere in reciprocal space is shown. The incident x-rays have wave vector k, and the diffracted rays have wave vector k'. The Ewald sphere has radius k = 27t/}t and its orientation is that of k relative to the origin 0 which is some reciprocal lattice point. The allowed diffiacted rays are those corresponding to the reciprocal lattice points which lie on the sphere. Since the electron 35 wavelength it is so small, the radius of the Ewald sphere is quite large relative to the reciprocal lattice; consequently, it may be regarded as a planar section through the reciprocal lattice. The resulting diffraction pattern then is just the intersection of this plane with the reciprocal lattice. 4.3.2 Single Crystal and Polycrystalline Diffraction Patterns The reciprocal lattice of a single crystal is just another lattice. For example, the reciprocal lattice of a bcc structure is an fcc lattice and vice versa. As a result, the intersection of a plane through the reciprocal lattice, and hence the diffraction pattern, consists of spots. On the other hand, a polycrystalline sample has a random distribution of orientations; therefore, its reciprocal lattice is a set of concentric spheres generated by moving the associated single crystal reciprocal lattice through all possible orientations about the origin. Consideration of only one rel-point7 constrained to move about the origin at a constant distance for all possible orientations reveals that the rel-point would trace out the surface of a sphere. All of the rel-points taken together would trace out a set of concentric spheres. Consequently, the diffraction pattern of a polycrystalline sample consists of rings since the intersection of a plane with the reciprocal lattice forms rings. As previously mentioned in the XRD chapter, however, the sputtered multilayers examined in this study are not polycrystalline—they have texture. 4.3.3 Textured Diffraction Patterns There are two types of texture important for multilayers: fibre and orientational. As previously mentioned in the XRD chapter, fibre texture refers to the alignment of one family of planes parallel to the plane of the layers. If a sample has fibre texture only, then 36 hu+kv+lw = [001] Figure 4-3: Portion of the reciprocal lattice for a (001) fibre-textured bcc film (after Tang and Thomass). its grains are randomly oriented about the fibre axis. If in addition, orientational texture is present, then the grains can possess only certain allowed orientations about the fibre axis. In this section, the expected diffraction patterns for such samples and their relevance to multilayers will be discussed. For additional details about this subject, the reader should consult Vainshtein9 or lReimer.IO For a sample of perfect crystallites with perfect fibre texture, the reciprocal lattice consists of stacked concentric rings—the orientation of a rel-point is now constrained by a constant distance from the fibre axis as well as a constant distance from the origin as shown in Figure 4-3. For the case where the beam axis is normal to the plane of the layers in a multilayer, i.e., the beam lies along the fibre axis, the resulting diffraction pattern will not contain all of the rings because the Ewald sphere intersects only those 37 rings in the zeroth layer. For example, in a perfectly [1 lO]-textured film, the 310, 321, and 420 rings will not be observed. For a sample of perfect crystal structure and perfect fibre and orientational texture, these concentric rings of the reciprocal lattice become several spots situated around the ring since only certain allowed orientations are present. The spacing and number of spots exhibit the symmetry of the orientational texture. For a beam of normal incidence, the diffraction pattern would consist of only these spots which lie on the rings of the zeroth layer. For the multilayers examined in this study, all of the rings allowed by Bragg's Law were observed just as in a polycrystalline sample. Does this indicate a lack of both types of texture? No, since most films do not exhibit perfect fibre texture, the fibre axis is distributed through some angle. Also, some small number of grains are probably randomly oriented, so the reciprocal lattice consists of concentric spherical belts and concentric spheres. So for an incident beam parallel to the normal of such a film, the resulting diffraction pattern will contain all of the rings observed for a random distribution.7 Likewise, most samples do not exhibit perfect orientational texture. The grains are distributed through some angle so the spots become arcs along the ring in the reciprocal lattice. The arc lengths are indicative of the strength of the orientational texture. The resulting diffraction pattern for a multilayer using a normally incident beam will consist of arcs of the rings observed in a random distribution. 38 4.4 Transmission Electron Microscopy The use of acronyms in electron microscopy is confusing at best. Two of the main modes of Operation in transmission electron microscopy have acronyms which refer to particular instruments—the transmission electron microscope (TEM) and the scanning transmission electron microscope (STEM). Unfortunately, several different instruments can be referred to using the same acronym. For example, STEM may describe a dedicated scanning electron microscope being operated in a transmissive mode, a dedicated transmission electron microscope being operated in a scanning mode, or a dedicated scanning transmission electron microscope. To avoid such ambiguities here, the use of TEM and STEM will be used primarily to refer to the two different modes of operation described in the following paragraph. In TEM mode, the beam cross section that reaches the sample (referred to as the probe size) has a diameter on the order of a few um, illuminates all of the region of interest simultaneously, and is very nearly parallel to the optical axis of the instrument column in the sample plane when operated in an imaging mode. By contrast, in STEM mode, the probe size is on the order of nm, and the beam scans back and forth across the much larger region of interest much like an electron gun scans across a television screen or a computer monitor. Just as in TEM mode, however, the beam is parallel to the optic axis in the specimen plane for the entire scan region as depicted in Figure 4-4.ll 4.4.1 Conventional TEM In this section, the primary components of a conventional TEM (see Figure 4-5) necessary for its use in TEM mode are described in descending order of their positions in 39 Scan coils Pre-field of the objective \/ Specimen / \ p ane Post-field of the objective Figure 4-4: STEM imaging mode (after Watt”). 40 l Filament —I:VJ Wehnelt cylinder I | I — Condenser aperture l I : Double condenser lens I | I ——> — <—— Sample _ — — — T — — — — Objective lens plane m m I l | l | I | l l l I _ — — — T — — — — Projection lens plane i l l l | Screen Figure 4-5: Schematic of the primary components of a typical conventional TEM. 41 the instrument column. These components must be well-aligned each time the microscope is used to minimize electron optical aberrations which are the same as those for visible light. Starting at the top of the column, the electron gun consists of a cathode filament (typically a hairpin), a Wehnelt cylinder, and an anode. The cathode filament is held at a large negative voltage and heated to emit electrons via thermionic emission. The Wehnelt cylinder is held at a slightly more negative voltage and helps to focus the electrons. A large potential drop between the grounded anode and the cathode accelerates the electrons. The condenser aperture collimates the electron beam, and the double condenser lens system demagnifies the beam onto the sample. The objective lens focuses the sample image. The objective aperture lies in the back focal plane of the objective lens and blocks out the diffracted rays to provide contrast when viewing the specimen. The SAD (selected area diffraction) aperture lies in the image plane of the objective lens and stops diffracted rays coming from outside the region of interest during diffraction, i.e., it "selects" the area of the sample contributing to the diffraction pattern. As this method of TED is typically the one used in most electron diffraction experiments, the procedure itself is often referred to as SAD. The intermediate lens focuses either the sample image or the diffraction pattern which is then magnified by the projection lenses onto either a fluorescent screen or photographic film. 4.4.2 Image and Diffraction Pattern Formation In this section, the formation of images and diffraction patterns in a conventional TEM operated in TEM mode will be discussed. Though the details of the processes may vary from one instrument to another, the basic concepts remain the same. 42 Object plane of the objective lens Sample Objective lens plane - — _ _ - Back focal plane of the objective lens - _ — _ — Objective aperture Object plane of the intermediate lens __________ a_—— ——-——— Intermediatelensplane Figure 4—6: The objective aperture blocks the diffracted rays thus providing contrast in the TEM image. \ Objective / /v I aperture Figure 4-7: Schematic of (a) BF- and (b) DF-imaging. I0 = primary beam, 1, = transmitted beam, Is = weakly scattered beam, and Id = diffracted beam. 43 Images obtain their contrast from the scattering of electrons in the sample. Though inelastic scattering may provide some weak contrast, the real contrast mechanism is diffraction. By inserting a small enough objective aperture, the diffracted beams are blocked from detection as in Figure 4-6. Consequently, regions that meet the Bragg condition appear darker than other regions. This difference results in differing contrast for crystallites and for differently oriented regions within a given crystallite such as a dislocation. This method of imaging in which the transmitted and weakly scattered electrons create the final image is known as bright-field (BF) imaging. In dark-field (DF) imaging, the incoming beam is tilted as shown in Figure 4-7 so that only one diffracted beam is used to form the image. In either case, the final image is that formed in the image plane of the objective lens. In a conventional TEM, this is accomplished by making the object plane of the intermediate lens lie coincident with the image plane of the objective lens as shown in Figure 4-8. By tracing all of the diffracted beams and the transmitted beam coming from the same spot on the specimen to their final destination, it is clear that the final image is of the sample and not the diffraction pattern. In order to view the diffraction pattern, the final image must be that of the image formed in the back focal plane of the objective lens. The SAD aperture is inserted and aligned such that the region of interest (R01) is imaged—only this region of the specimen will contribute to the diffraction pattern. The ROI size is given by D/M where D is the diameter of the SAD aperture and M is the magnification of the objective lens. So for D = 40pm and a typical value of M = 40, the ROI size would be 1m. Below this size, electrons from outside the ROI would contribute to the diffraction pattern due to spherical aberration of the objective lens. Upon switching to diffraction mode in a conventional Object plane of the objective lens Objective lens plane Back focal plane of the objective lens Image plane of the objective lens Object plane of the projector lens Projector lens plane Image plane of the projector lens 44 Sample Objective aperture Object plane of the intermediate lens Intermediate lens plane Image plane of the intermediate lens Figure 4-8: Image formation in a conventional TEM. The rays for the diffracted beams are blocked beyond the plane of the objective aperture but are included here for illustration (after Crimp”). Object plane of the objective lens Objective lens plane - Back focal plane of the objective lens Image plane of the objective lens Object plane of the projector lens Projector lens plane Image plane of the projector lens 45 Sample Object plane of the intermediate lens SAD aperture Intermediate lens plane Image plane of the intermediate lens Figure 4-9: Diffraction pattern formation in a conventional TEM (after Crimp”). 46 TEM, the object plane of the intermediate lens is made coincident with the back focal plane of the objective lens so the final image is that of the diffraction pattern as shown in Figure 4-9. By tracing all of the beams diffracted in the same direction from the three spots at the specimen which combine to make one spot on the final image, it is obvious that the final image is that of the diffraction pattern. Also, the objective aperture should be removed during operation in diffraction mode to observe all of the diffraction pattern possible. 4.4.3 VG HB501 FESTEM The VG HB501 FESTEM is a dedicated field-emission scanning transmission electron microscope manufactured by VG Instruments and is the EM used in this study. The term "field-emission" refers to the type of cathode used by the FESTEM in which the electrons are extracted by a strong electric field instead of by thermionic emission. A field-emission gun forms a brighter and smaller probe, but it can only be used in ultra- high vacuum (UHV) on the order of lO'Htorr which greatly increases the cost and maintenance of the instrument. In the remainder of this section, only the primary components of the FESTEM necessary for imaging and for observing diffraction patterns are discussed; however, the FESTEM includes several other components used in other modes of operation. As shown in the schematic diagram of Figure 4-10, the entire column is upside down relative to a conventional TEM. This design allows the sample stage to be mounted rigidly on top of the objective lens to provide greater mechanical and thermal stability required for high magnification and simplifies detector mounting.15 47 BF detector I Collector aperture Annular DF detector Sample — — — — —L — — — — Objective lens plane I I __ _ <— ROA — <— SAD aperture Scanning coils EIZ I I I I I I I I Optic | axis l I I l I I I I I Double condenser lens — VOA I‘A_I Field-emission gun I I Figure 4-10: Schematic diagram of the primary components used in TED for the VG HB501 FESTEM. 48 Starting with the field-emission electron gun at the bottom of the column and working upward, the gun is followed by the virtual objective aperture (VOA). It is used instead of the real objective aperture (ROA) during certain operations of the microscope to prevent detection of unwanted x-rays when the x-ray detector, which is very close to the RCA, is in use. As in the TEM, the double condenser lens system controls the demagnification of the source. It is followed by the scanning coils which scan the beam back and forth across the sample. The SAD aperture, the RCA (corresponding to the objective aperture in the TEM), and the objective lens serve basically the same purposes here as in the TEM. The specimen is situated very near the center of the objective lens. The detection of the electrons is performed quite differently compared to the TEM. The diffracted electrons are collected by the annular dark-field detector consisting of a scintillator, a light guide (since the detector is not mounted along the column axis), and a photomultiplier (PM) tube. This detector collects the electrons scattered into an annular cone in the approximate half angle range 0.04 rad to 0.2 rad16—-—these are the angles at the specimen before the beam is compressed by the post specimen field of the objective lens. A hole through the scintillator and light guide allows the transmitted and weakly scattered electrons to pass through the annular dark-field detector, then the collector aperture, and finally into the bright field detector which also consists of a scintillator and PM tube. The signals fiom these detectors may be viewed on two oscilloscope screens so that a BF and a pseudo-DF image of the sample can be viewed simultaneously. The BF image is formed in the same manner as it is in a conventional TEM; however, since all of the diffracted beams are used to form the "DF image" instead of only one, it is referred to as a pseudo-DF image. 49 Since the electrons are detected electronically, the diffraction patterns can be displayed on an oscilloscope screen in several different ways. One of the most useful is to scan across the rings of the diffraction pattern in a line to display an intensity profile of the rings. Since this type of display looks very similar to an XRD diffraction pattern recorded by a diffractometer, often the allowed rings will be termed allowed lines. Also, a display on one of the two viewing screens can be imaged on a third screen and recorded using an instant black-and-white Polaroid camera. 4.5 Sample Preparation and Typical Settings of the FESTEM The samples used for TED had to be sputtered thinner than those used for the other studies. Typical total thicknesses were chosen to be near 500A to limit the effects of chromatic aberration. For larger thicknesses, the energy lost by the diffracted electrons becomes large enough to result in serious focusing errors, and for still greater thicknesses, the sample simply becomes opaque to electrons. Either NaCl or KCl were chosen as substrates for the sputtered samples since they dissolve in water unlike the Si and A1203 substrates. For a given sample, the salt substrate was cleaved to the required size (around 2-4 mmz) using a new razor blade for each sample run, and the multilayer was floated off the substrate in distilled water onto a Cu grid mesh. The grids used were 3mm in diameter, about 15pm thick, and had 200 lines to the inch in both directions of the square mesh.17 They had a shiny finish on one side and a matte finish on the other. The sample was always placed on the matte side to aid the electrostatic attraction between the sample and the grid. Of course, the sample has to be viewed between the grid bars since the grids are opaque to electrons. 50 The VG HB501 FESTEM was operated at 100 kV using the cold stage for the Fe/Si and Fe/{FeSi} samples which kept them at approximately -140°C when not interacting with the beam to minimize interdiffusion.‘8 At least 3 different regions of the sample were investigated to verify that the samples were homogeneous. A region was selected, the column aligned, and the diffraction patterns recorded at two different camera lengths—typically 0.5m and 1.0m. Typical aperture sizes used were 300 um for the VOA aperture, 100 pm for the SAD aperture, and 50 um for the collector aperture. Since the RCA aperture was not used during TED measurements, its size is unimportant. 4.6 Analysis 4.6.1 Structural Determination In the multilayer TED patterns, typically many lines were observed allowing an unambiguous determination of the average structure of the bilayer unit or at least one of the component layers—in the Co/X multilayers, no lines uniquely associated with Co were observed with the VG HB501 FESTEM. Unlike XRD, the d-spacings cannot be determined unless a calibration standard is used to determine the diffraction constant XL. In a conventional TEM, AL can be measured accurately once for a given setting and thus be determined for all samples. Only one calibration is required since each sample region can be placed at the exact same height in the microscope by using a eucentric stage to hold the sample. Such a stage can be tilted while observing the R01 and changing the height of the specimen. When the sample image movement is minimized, then the R01 coincides with the tilt axis and is placed at a specific height in the column. The other requirement for making only one calibration is that the high voltage transformers are 51 quite stable so the electron wavelength does not change. Unfortunately the VG H3501 does not have a eucentric stage; therefore, a calibration involving simultaneous examination of both the sample and a suitable standard would have to be performed for every measurement if it were desired to determine d-spacings accurately. Structural identification can be done without such a calibration since AL is constant for a given ROI. Consequently, the ratio of one ring radius to another is independent of XL, and the allowed lines may be determined. This method is known as the method of ratios. In examining the modified Bragg equation, Rd = AL, it is clear that the ring radii are inversely pr0portional to the d-spacings; consequently, Rl/R2 = d2/d1 independent of the diffraction constant. The actual procedure consists of measuring the ring radii and calculating the ratios of each ring radius to the first ring of significant intensity (the dominant line) which is almost always the first allowed line of the given structure. Comparison of these ratios to the predicted values for the various structure types allows identification of the structure to be made. 4.6.2 Compound Identification On occasion, additional lines may be observed which do not correspond to any ratios for the determined structure. Ifthe structure is not fcc or bcc, then double diffraction may have led to the presence of forbidden lines; otherwise, compounds may have formed at the interfaces from the constituent components resulting in additional Bragg lines. In either case, some assumptions must be made about the d-spacing of the dominant line. In determining the possibility of a forbidden line, the expected d-spacing should be determined by using the relationship between planar spacings, lattice constants, and 52 Miller indices for a given structure. The expected ring ratio can then be calculated as in the preceding section and compared to the actual value. In determining the possibility of compound formation, the Powder Diffraction File may be consulted to find possible equilibrium compounds that may be formed from the constituent materials. The ring ratios for the listed d-spacings can be found as before and compared to the actual values. Since typically in this study only one or two additional lines were ever observed, the strongest statement that may be made is that the line(s) are consistent with the presence of certain compounds. lManfred von Heimendahl, Electron Microscopy of Materials: An Introduction, (Academic Press, San Diego, 1980), p. 96. 2Manfred von Heimendahl, ibid. 3Ian M. Watt, The Principles and Practice of Electron Microscopy, (Cambridge University Press, New York, 1985). 4Manfred von Heimendahl, ibid. 5 Thomas and Goringe, Transmission Electron Microscoy of Materials, (J. Wiley & Sons, New York, 1979). 6Hirsch, Howie, Nicholson, Pashley, and Whelan, Electron Microscopy of Thin Crystals, (Krieger Publishing, 1977). 7Quantities such as points or planes associated with the reciprocal lattice are denoted by "rel". 81.. Tang and Gareth Thomas, J. Appl. Phys. 74 (s), 5025 (1993). 9Vainshtein, Structure Analysis by Electron Diflraction (Pergamon, New York, 1964). 10Ludwig Reimer, Transmission Electron Microscopy: Physics of Image Formation and Microanalysis, (Springer-Verlag, Berlin, 1984). ”Ian M. Watt, ibid., p. 71. 12Ian M. Watt, ibid. 53 13 M. A. Crimp, course notes, Michigan State University (1994). 14M. A. Crimp, ibid. l5Ian M. Watt, ibid., p. 75. l6VG HBSOl operation manual. 17Ian M. Watt, ibid., p. 82. 18Ludwig Reimer, ibid., p. 440. CHAPTER 5 X-RAY ABSORPTION FINE STRUCTURE 5.1 Theory X-ray Absorption Fine Structure (XAF S) refers to the oscillations in the x-ray absorption in a material which occur beyond a given absorption edge as a function of the incident photon energy. The defining equation for this absorption in a transmission experiment is Izloe’”x (Eqn.5-1) where 10 and I are the intensities of the incident and transmitted x-ray beams respectively, p. is the absorption coefficient, and x is the sample thickness. A marked increase in the absorption occurs at an absorption edge where the incoming x-ray energy just matches the binding energy of a core electron. If the atom containing this electron is isolated, then the absorption decreases monotonically; however, if it is surrounded by other atoms, fine structure variations in the absorption occur known as XAF S. Until recently, these oscillations were divided into two different regions. In the region approximately 40-1000 eV above the edge where weak single scattering dominates, the resulting oscillations were referred to as EXAF S (Extended X-Ray Absorption Fine Structure). In this region, the Born approximation is valid, i.e., the assumption that the wave states of the ejected photoelectron are plane waves does not cause significant error—the next section will describe this process in greater detail. 54 55 Before this EXAFS region, strong multiple scattering dominates since the photoelectron is ejected with little kinetic energy. This much smaller region is the XANES (X-Ray Absorption Near Edge Structure) region. Here the Born approximation introduces unacceptable error, and until recently, made analysis quite difficult. Now that improvements in calculating the XANES region theoretically have been made, it is becoming more customary to combine the two regions into one—the XAF S region. 5.1.1 Physical Origin1 The probability that a photon will be absorbed by a core electron depends upon the initial and final states of that electron. The initial state is just the wavefunction of an electron in the core of the atom. For an isolated atom, the final state of the ejected photoelectron is represented by an outgoing spherical wave leaving the absorbing atom. If there are surrounding atoms, then the final state is represented by a sum of this outgoing wave and all the backseattered waves from these neighboring atoms. It is the interference between these waves that give rise to the oscillations in the absorption coefficient. Proper analysis of these oscillations provides structural information about the local environment (within ~ 6A) around a particular type of atom. 5.1.2 The XAFS Equation The following equation is valid for an unoriented sample incorporating Gaussian disorder and many-body effects.2 x (k) = Z-gz—Sflkmat) exp( — 218s?) exp[— 2(Ri — Amt] sin[2kRi + 5i(k)] (Eqn. 5-2) 1 1 56 The term sin[2kRi + 6,(k)]/(k2R2) is the functional form of the outgoing spherical wave with an amplitude of F1 (k) = mti (2k)/21r hzk where k = wave number of the outgoing spherical wave, m = electron mass, ti(2k) = factor proportional to the backward scattering amplitude of the outgoing wave which is characteristic of the backseattering atom, N, = number of atoms of the type "i" in a given shell, and R, = average distance from the center of the ith shell. The term S3 (k) is the amplitude reduction factor associated with the relaxation of the wavefunctions of the "passive" electrons, i.e., those electrons which were not ejected. Pi> electron in the central atom and p,’ is the relaxed state of the ith passive electron. The 2 where p, is the wave function of the ith passive It is determined by SE) = MI crystal monochromator is used which has a resolution of better than 2 eV near 6.5 keV. One reason for the choice of two crystals is that by detuning them, the fimdamental to harmonic ratio is improved. Direct tests were performed to check for the presence of harmonics in every experiment. 5.2.2 XAF S Measurement Techniques Several different techniques may be used to measure the XAF S. Either x-ray or electron detection may be used. The direct measurement of the transmitted x-rays is the most straightforward technique. In it, the intensity of the x-rays incident upon the sample (10) and through the sample (I) are measured—the absorption is then given by ux = ln(Io/I) from Equation 5-1. Other measurement techniques monitor processes that are proportional to the absorption including x-ray fluorescence and Auger electron emission. In fluorescence measurements, the intensity of the fluorescent x-rays (If) is measured along with 10. These fluorescent x-rays are generated when electrons from higher shells fill the hole left behind by the ejected photoelectron resulting in a characteristic photon emission. Auger electron detection is one of several electron detection methods including total electron yield. Auger electrons are emitted when the emitted fluorescent photon described above is absorbed by another electron in the same atom causing its ejection— this ejected electron is an Auger electron. The choice of these techniques for a XAF S measurement depends upon the sample being examined. In this study, the fluorescence technique was chosen over transmission for the following reasons. One, the samples would require either sputtering onto transparent substrates such as kapton tape or removal from their A1203 or Si substrates. Two, for the 6O thickness range of samples studied here, the systematic error corrections from thickness effects are negligible for fluorescence measurements but not for transmission measurements. These "thickness effects" occur whenever a portion of the incident beam is not attenuated by the sample in the same manner as the remaining incident beam—this discrepancy can result from anything such as pinholes in the sample to disparate energies in the beam (harmonics)—the "leakage" becomes a greater portion of the total signal as the sample thickness increases. As a result the absorption signal appears to depend upon the sample thickness resulting in systematic errors in the measured absorption.5 The reason for choosing fluorescence over electron detection is not as obvious. For those elements with Z < 30 including Fe, Co, and Cu studied here, the Auger channel for K-edges dominates the fluorescence channel;6 however, the origination of the Auger electrons is limited to within approximately 30A of the surface.7 Perhaps as a result of this surface limitation, attempted electron-yield measurements had a very poor signal-to- noise ratio, so the fluorescence technique was chosen. In fluorescence, the measured signal u'(E) is proportional to the absorption as shown in the following equation8 for thick samples: ' If p.(E)sin9 11 (E) =- ~ . . I uT(E)/sm9+uT(Ef)/srncp 0 (Eqn. 5-3) where 9 is the entrance angle of the incident x-rays, (p is the exit angle of the fluorescent x-rays, 11r(E) is the total absorption coefficient at energy E, Ef is the fluorescent energy, and the equation is integrated over the angles (p subtended by the detector. In the studies performed here, an in-house Lytle detector9 operating in fluorescence l 11, I - — <— Soller slits filter / ionization chamber Figure 5-1: Schematic diagram of a Lytle detector. mode was used. A schematic diagram of this detector is shown in Figure 5-1. The incident beam (10) enters one end through a slit and strikes the sample. The transmitted beam (I) then exits the detector at the opposite end where another detector may be positioned if it is desired to measure the transmission signal. The majority of the fluorescent x-rays emerge from the sample at a 45° angle as shown in the diagram. The x-rays then pass through a filter selected such that the elemental composition is typically one less in atomic number than that of the absorbing atom, e.g., Mn is used as the filter for Co. The filter reduces the intensity of x-rays with energies above the absorption edge energy. After passing through a set of Soller slits, the fluorescent x-rays then enter the ionization chamber which is typically filled with flowing Argon to enhance the ionization process. Concerns about the presence of harmonics were sometimes handled by using static air in this chamber causing the detected harmonic to fimdamental ratio to decrease since the air is not as absorbing as the Ar. The Io detector had either static air or a mixture of N2 and He flowing through it. Though a rough rule of thumb for determining 62 the gas mixture used in this detector is to decrease the intensity of the primary beam by 10%, ultimately, the choice is made by empirically determining what provides the best signal-to-noise ratio. 5.3 MacXAFS Analysis All analysis of XAF S data was limited to the EXAF S region and was performed using MacXAF S v. 2.0 or v. 3.6 both of which are based on the University of Washington EXAFS analysis package written by Bruce Bunker. Subsequent generations of this software have been written with many different contributors, and the package has been ported to several different machine formats. The MacIntosh format modifications used in this study were made by Charles Bouldin of the National Institute of Standards (NIST) and Lars Furenlid of the NSLS. This software package enables the user to perform all the required steps of an analysis. First, the raw data is converted into the format required by MacXAF S. An example of the raw data from a fluorescence measurement of a 2000A sputtered Co film is given in Figure 5-2. Second, the data are normalized to the step height so that samples with different total numbers of absorbing atoms—different thicknesses in this study—may be directly compared, and an appropriate background is subtracted to isolate the EXAFS oscillations. At this point, the data has the form of Equation 5-2 which is x(k). An example of this CHI data is shown in Figure 5-3. Since R, is different for each shell, the EXAFS corresponding to different coordination shells oscillate at different frequencies. As a result, the third step consists of Fourier transforming these oscillations in k-space into a more physically meaningful form—the radial distribution function (RDF). The 63 XMU __,J O r _ . 1 . 1 1114 t l r 11 Iml I I‘LLMLLI 7.5 7.6 7.7 7.8 7.9 8.0 8.1 8.2 8.3 8.4 8.5 8.6 8.7 Energy (keV) Figure 5-2: Raw data from a fluorescence measurement of 3 2000A sputtered Co film. 0.10 . . - T . I . . . , , , 0.08 L «I 0.06 2 . l 0.04 L I L 1 0.02 - 0.00 . AVAVAVAX .._\/ III A CHI l I- -0.04 I -0.06 a 2 4 6 8 10 12 14 16 k (5.") h— >— -0.08 Figure 5-3: CHI data for the same 2000A sputtered Co film shown in Figure 5-2. 64 4O ' I ' l I I T l ' l I f f I RSP l r l l L A l 1 l .1 l 1 l R (A) Figure 5-4: RDF for the same 2000A sputtered Co film shown in Figure 5-2. RDF for the 2000A sputtered Co film is shown in Figure 5-4. The RDF consists of peaks which may or may not be resolved corresponding to the various shells surrounding the central atom. Integration of this function provides the number of atoms contained within the limits of integration. Unfortunately, this analysis does not provide the actual RDF. Since the phase shift 5,(k) of each backscattered wave is not constant but has a k- dependence, the Fourier transform of x(k) results in a pseudo-RDF. Additional discrepancies between the Fourier transform and the actual RDF result fiom the finite range of integration as well as an improper background subtraction. Consequently, the pseudo-RDF cannot be analyzed directly to obtain the coordination number for each shell. Instead, a peak in the RDF is backtransforrned to isolate the contribution of that coordination shell, and a non-linear least squares fit using the Levenberg/Marquardt search strategy is made to an appropriate experimental standard. In this manner, the 65 coordination number (N), the interatomic distance (R), and the Debye-Waller factor (0) are determined for the given shell(s). The standard and the sample must be treated in the exact same manner when performing the data analysis to obtain meaningful results. MacXAFS v. 3.6 uses an interactive graphical interface which allows the user to treat the sample and the experimental standard exactly the same more easily than previous versions. The user invokes this interface by selecting the graph/hand icon located near the center of the given XMU, CHI, RSP, or ENV display. In the following sections, the specific details of each step are discussed. 5.3.1 Data Conversion (Raw —> XMU Files) First, the raw data is converted into a format readable by MacXAFS—an XMU file. Several beamline formats are included in MacXAFS v. 3.6; however, any data format can be defined by the user. Once this conversion is complete, the background is removed creating a CHI file. 5.3.2 Normalization and Background Removal (XMU —-) CHI Files) Prior to subtracting the background, any single point glitches which exist in the region to be examined are removed since they may affect what background is subtracted. Though MacXAFS includes a method for the removal of wider multiple point glitches, they should be dealt with during the course of the experiment. Oftentimes, such glitches are caused by the underlying sample substrate, so if the sample is rotated, they will disappear or move allowing acceptable regions of different scans to be combined. In any case, the glitches must be removed since Fourier transformations are performed later causing significant broadening of the glitches. 66 Ideally, the CHI data would be normalized according to x(E) = [u(E) - I1(,(E)]/p0 where u is the absorption coefficient and no is the bare atom absorption. Unfortunately, no cannot be measured independently nor determined theoretically near an absorption edge; therefore, the following normalization is used: = ME) - ME) ME.) x(E) (Eqn- 5-4) where u(Eo) is the value of p at the absorption edge, i.e., the step height, and pb(E) is a suitable background since the actual background cannot be measured. The data are normalized to the step height since different samples have different total numbers of absorbing atoms. Since the actual background cannot be removed, a suitable background consisting of a linear fit to the pre-edge region and a cubic spline fit for the remaining data is used. The user must be careful to insure that the cubic spline does not follow the EXAFS oscillations. This background removal is satisfactory as long as both the experimental standard and the sample have similar backgrounds subtracted. The remainder of this section discusses the specifics of this procedure. Using Interactive Data Plot (IDP), the various scans for a given sample are simultaneously displayed to determine if each background should be subtracted separately. Though the article by Vaarkamp, et a1.10 suggests that backgrounds should always be removed separately for each scan to average out statistical errors during the subtraction process, no statistically meaningful differences in the parameters obtained from the analysis were observed in this study unless the scans had visibly different 67 backgrounds. Dissirnilar backgrounds sometimes occur when the gas mixture in detection chambers has not yet come to equilibrium. If the backgrounds do not need to be removed separately, then the XMU files created after the data conversion are combined using FMERGE which averages the data from all the scans included in the merge. Since the statistics are better for this combined data, the background removal may actually be superior to those performed separately. In determining the pre-edge background subtraction, two points, EFl and EF2, are chosen to give the best linear fit to the pre-edge. Obviously, as large a range as possible is selected while avoiding substrate and monochromator glitches. In determining the cubic spline portion of the background, the fewest number of knots (zero crossings) which allow for a suitable subtraction are used. The best first attempt is one knot with an unspecified position. Also, the data should be weighted more heavily toward the high k- range since low-k data tend to be less reliable than high-k data.11 Typically, a fit weight of three is used for most metals. The range of data selected includes a minimum energy no less than 20 eV above the edge and preferably higher. The maximum energy is determined by the signal-to-noise ratio or by the presence of glitches either of which could significantly alter the subtracted background. Finally, an easily observable feature of the edge step is chosen to specify the edge energy—consistency is much more important than choosing an exactly correct value. This choice is determined by selecting the zoom option which displays the derivative of the step so the point of inflection is easily seen. The edge step is selected such that its bottom is a good average of the EFI and EF2 points. The top of the step is then chosen 68 such that the background is approximately midway between the extrema of the oscillations. The selected background can be previewed visually before actually removing it by selecting the "poly-fit" button after the parameters have been chosen. Once the background is removed, Tekplot is used to examine its derivative. If it seems to follow the oscillations of the raw data at all, then the parameters must be adjusted and the process repeated until this is no longer the case. 5.3.3 Fourier Transform (CI-II —> RSP Files) If the backgrounds were removed separately for each individual scan, then these CHI files are merged. Now only one CHI file exists to be Fourier transformed. Since the CHI data extend over a finite range of k, the Fourier transform introduces Fourier filtering errors; therefore, the width of this range must be the same for all samples and standards so the errors will be the same for all. To reduce the effects of this Fourier filtering, the minimum and maximum k-values are chosen at zeroes in the CHI data. Once this k-range window is set, it can be moved without changing its size. The largest k-range possible is selected though it is more important to determine a zero accurately than to increase the size of the k-range. Again, the k-weighting chosen will probably be three; however, the CHI data should verify this choice. The value of the k-weighting is chosen such that the weighted CHI plot (k"x(k) vs. k) is closest to having equal amplitude over the entire range. This check can be made by selecting Wgt'd Chi Plot in the Plot Options during the background removal. Further reductions in Fourier filtering errors are made through the use of a Hanning function or window W(k). By combining the data with W(k), the data are not as sharply 69 truncated as they would be with a step function thus minimizing the introduction of spurious peaks into the pseudo-RDF. Though this method minimizes truncation ripple, it does cause some peak broadening. The values of dkl and dk2 determine the width of this Hanning window which is defined by the following: W(k) = sinzIn(:(_d:3m)I kmin < k < (kmin +dk1) =1 (kmin +dk1) ENV Files) After obtaining the pseudo-RDF, a given peak is backtransformed to isolate the contribution of a specific coordination shell. In this study, only the nearest neighbor (NN) peak was backtransformed except in the case of Fe where the first two shells are simultaneously backtransformed since the two peaks in the pseudo-RDF are unresolved— these two peaks are unresolved for all bcc structures. The NN peak is typically the easiest to analyze since multiple scattering and the k-dependence of the electron mean-free path are minimized for it. Just as in the case of Fourier transforming, all samples to be compared along with the standard must use the same backtransform range. Once the window has been set for the first sample, it can be moved without changing its width. In choosing the backtransform range, it is absolutely imperative to include all of the peak— 70 even the tails. In a comparison of Fourier filtered and unfiltered spectra created using FEFF--a theoretical XAF S modeling software program which does not use the Born approximation making it the best generator of theoretical XAF S data currently available, 1.12 determined that omitting even a small portion of the main peak results Vaarkamp, et a in relatively large errors. Comparisons made in this study show that the coordination number obtained through the fitting procedure decreases significantly as the backtransform range is decreased even though both the standard and the sample are backtransformed over the same range width. 5.3.5 Least-squares Fit Once the contribution to x(k) by the first shell has been isolated, i.e., the backtransform has been completed, several parameters are varied to obtain the best agreement between the standard and the sample. The maximum number of parameters 11 which may be fit are determined by n = 2AkAR/7t where Ak is the momentum space range and AR is the real space range of the first shell.13 The coordination number N, the nearest-neighbor distance R, and the Debye-Waller factor 6 (denoted by SS in MacXAFS) are typically varied in the fitting procedure. The data fitting weight is usually three since the high-k data are more reliable. Typically, o is set to zero for the standard since it is unknown; therefore, the value obtained for the sample is a relative one. Also, as a general rule of thumb, AR z 0.01-0.02A and AN/N z 15-20%. Finally, the goodness of fit parameter should be significantly less than one. 71 lB.K. Teo and DC. Joy, EXAFS Spectroscopy: Techniques and Applications, (Plenum Press, New York, 1981), pp. 13-17. 2D.C. Koningsberger and R. Prins, X-Ray Absorption: Principles, Applications, Techniques of EXAFS, SEXAFS, and XANES, (John Wiley & Sons, New York, 1988), p. 45. 313.x. Teo and DC. Joy, ibid., p. 36. 4DC. Koningsberger and R. Prins, ibid., p. 127. 51bid., p. 108. °Ibid., p. 455. 7Ibid., p. 93. 81bid., p. 92. 9F.W. Lytle, R. Greegor, D. Sandstrom, E. Margues, J. Wong, C. Spiro, G. Huffman, and F. Huggins, Nucl. Instr. Meth. A226, 542 (1984). 10M. Vaarkamp, I. Dring, R.J. Oldman, E.A. Stern, and DC. Koningsberger, Phys. Rev. B 50(11), 7872 (1994). llD.C. Koningsberger, ibid., p. 232. 12M.Vaar1tamp, ibid., p. 7874. l3F.W. Lytle, D.E. Sayers, and EA. Stern, Physica B 158, 701 (1989). CHAPTER 6 CO/X MULTILAYERS 6.1 Introduction Metallic multilayers formed by alternating layers of ferromagnetic metals with layers of non-ferromagnetic metals exhibit a number of striking macroscopic magnetic properties. A loss of ferromagnetisml or a rapid large increase in the magnetoresistance2 as the layers become thinner are two distinctly different examples of such properties. The latter behavior is known as giant magnetoresistance (GMR). Although the macroscopic properties are commonly reported as a function of layer thickness, two XAF S studies of Cu/F e multilayers illustrate the inadequacy of attempting to understand these macroscopic properties without better structural information. Pizzini, et a1. studied samples having 15A Fe layers with a variety of Cu layer thicknesses.3 Their XAFS data indicated a structural transition occurred as the Cu layer thickness varied. Below 3 layer thickness of 10A the Cu layers were bcc and as the layer thickness increased a transition to a fcc structure occurred. Not surprisingly, an increase in local disorder was observed in the thickness regime between the limiting bee and fcc structures. This thickness regime was correlated with changes observed in the GMR. Cheng, et al. studied a series of Cu/Fe samples for which the Cu layer thickness was dominant.4 Their XAF S data identified a structural transition in the Fe layer that was correlated with the observed loss of magnetism in their samples. Many Co/X systems show similar behaviors in their macroscopic magnetic properties but obtaining reliable structural information can be more difficult. 72 73 Complications arise because of the difficulty in distinguishing between hop and fcc structure for thin Co layers. In this chapter, three Co/X systems will be discussed— Co/Ag, Co/Cu, and Co/Mo. For thin Co layers, both Co/Ag and Co/Cu exhibit GMR, while Co/Mo displays a loss of ferromagnetism. Aside from these interesting magnetic properties, these systems provide three interesting cases for comparison. The fcc structures of Ag and Cu are very similar to the hcp structure of Co; however, Ag (d<111> = 2.359A) has a rather large lattice mismatch with C0 (d<002> = 2.03 5A) while the mismatch of Cu (d<||]> = 2.087A) with Co is small. On the other hand, Mo has a bee structure which is quite different from that of Co. In this chapter, the results of an earlier study of Co/Ag will be summarized, and the results of the other two Co/X systems will be presented in detail. 6.2 Co/Ag In this section, the results of an earlier publications are summarized for comparison with the other two Co/X systems. In this study, most of the samples had a 35A Ag layer and a C0 layer varying in thickness from 6-35A. The XRD data were consistent with a strong <111> texture and good layering. All of the samples except one had a saturation magnetization that was 85% or more of the measured saturation for a 5000A Co film. This Co film's magnetization was 1390 emu/cm3 which is within 4% of the value for bulk Co. No clear dependence upon layer thickness was observed. In the XAF S portion of the study, the Co 35A/Ag 35A sample was used as the standard, and R, N, and (I2 were determined for the remaining samples. The results of these fits along with the magnetization data are presented in Table 6-1. The only significant change is in the 74 Table 6-1: Saturation magnetization and fitted EXAFS results.5 N values are estimated to be accurate to within i 15%. GMR values are current-in-plane data at 5K from PA. Schroederé—the percent change is relative to the resistance at or near magnetic saturation. Sample M53, GMR Co A/Ag A N R(A) 62(A'2) (103 emu/cm3) (%) Pure Co 12 2.502 1.446 35/35 12 2.502 0 1.28 14 20/35 10.8 2.50 -0.0006 0.137 1.41 20 16/35 10.3 2.51 -0.0004 0.100 1.08 22 12/35 9.7 2.50 0.0007 0.194 1.25 24 6/35 8.4 2.49 0.0009 0.053 1.33 32 6/6 7.9 2.50 0.0008 0.007 1.26 ~l 4/ 12 Poor Fit 1.18 >20 4/15 Poor Fit 1.20 60 4/25 No data 1.35 40 coordination number which varies from 7.9-12 and gives a simple scaling with Co thickness. Since for this same thickness range, both the GMR and the manner in which the magnetization approached saturation changed dramatically, the final conclusion was that no simple correlations between the local structure and the magnetic behavior of these multilayers exist. 6.3 Co/Cu Cu/Co multilayers exhibit GMR7'8 and have been the subject of a number of structural studies. Lamelas, et al. used RHEED complemented by demanding in-plane XRD studies and found that the Co in Cu/Co prepared by molecular beam epitaxy (MBE) was stabilized in a fcc structure for Co layer thicknesses up to 20A and that interfacial 75 coherence persisted up to 40A of Co.9 Despite the poor x-ray contrast between Cu and Co, they were able to identify an expansion of > 1% in the Co lattice. Several early NMR studies claim the Co in Cu/Co has a mixture of fcc and hcp environments, and the proportions of the two structures appear to depend upon the method of sample preparation.‘0" "'2 These studies do agree that the Cu/Co interfaces in both MBE and DC sputtered samples with each layer thickness below 40A can be abrupt. The scale for interfaces is monolayers (ML) and the value is one to a few. More recent NMR studies of Cu/Col‘l"l4 introduce some additional preparation parameters but do not alter the conclusion that thin Cu/Co samples prepared by DC sputtering and MBE can have abrupt interfaces. XAFS can provide valuable structural information about the local environment. Pizzini, et a1. did XAFS studies on 5 thin, evaporated Cu/Co samples and concluded that the atomic layer spacings perpendicular to the layers remained that of bulk Cu and Co, but in the layers, Cu contracted and Co expanded.15 Unfortunately, the contrast between Co and Cu scattering functions was too weak to allow identification of the fraction of Co and Cu scatterers in the first shells surrounding each element. In this section, the XAF S results for DC sputtered Cu/Co samples are presented. These results are complemented by AXRD data where AXRD denotes XRD studies in which the x-ray energy has been tuned just below the Co K edge and thus the anomalous dispersion enhances the x-ray contrast between Cu and Co. 76 6.3.1 Sample Fabrication The samples were prepared using the previously described sputtering system. The substrates were 1 cm squares cut from a Si wafer. The total thickness of each sample was typically 60 to 70 times the basic bilayer unit. XAF S studies were done on a series of Co 15A/ Cu xA samples where x = 9, 16, 25, and 35. This series was chosen for three reasons: (i) this set of samples has layer dimensions near extrema of anti-ferromagnetic and ferromagnetic couplings, (ii) bilayer and individual layer thicknesses are consistent with NMR reports of sharp interfaces, and (iii) the two samples with smaller x values have thicknesses comparable to the evaporated samples studied by Pizzini, et al.15 A survey AXRD study of Co/X samples included the Co 15A/ Cu 9 and 16A samples, a Co 15A] Cu 9.5A with buffer layers, and a C0 4A/ Cu 25A sample. 6.3.2 XAF S Results XAF S studies were done on line X23A2 at NSLS described in section 5.2. Data for the Cu and Co K edges were taken using a Lytle detector operating in a fluorescence mode. The samples were mounted at 45 degrees to the beam and px = 3 filters were placed before the Soller slits. A Ni filter was used for Cu data, and a Mn filter was used for Co data. The incoming beam was monitored by an ionization detector filled with static air rather than a more absorbing mixture to reduce complications from harmonic contributions. A XAF S scan of Co for a Co/Ag sample directly confirmed the lack of harmonic contamination; there was no evidence of the Ag K edge which would be produced by a third harmonic. This gas selection did not compromise sensitivity. The same Co/Ag sample had been part of an earlier Co/Ag study on the harmonic free line 77 X23B where the Lytle detector16 was filled with flowing Ar. Analysis of the present Co data relative to a pure Co standard gave results identical to those of the earlier study. All the present data were taken at room temperature. Typically 5 scans were performed for most samples, and 7-9 scans were done for those samples with weaker signals. Results from different scans were combined during the analysis using MacXAFS v. 2.0. Data from pure Co and Cu foils were used as the standards in this analysis. Since comparison of the present XAF S results for DC sputtered samples with the earlier results for evaporated samples was a major goal of this experiment, the XAFS results will be presented in a manner permitting direct comparison to those results. The oscillations for Cu typically extended to a zero crossing near 11.5A'l and occasionally further, but the oscillations for Co generally became noisy above a zero crossing near 10A". As long as the same momentum range was used for both the standard and unknown in forming the Fourier transforms used to isolate the first shells, the results from fitting the first shell back transforms were independent of the actual momentum range used. Given this independence, the Cu and Co data were fit over comparable momentum ranges. Table 6-2 presents the detailed XAF S results under two different constraints. First, the number of nearest neighbors were varied, and second, the coordination number was fixed at 12. The latter condition was used by Pizzini, et a1.15 and was supported by their model calculation which indicated that nearest neighbor Cu and Co atoms were indistinguishable. Paralleling their analysis, the present study used experimental data for pure Cu to fit Cu data and experimental data for pure Co to fit Co data. Despite the somewhat more limited momentum range, the quantity 2AkAR/7t is nearly 5 and thus fitting 2 or 3 parameters is valid. 78 Table 6-2: EXAF S results for Co/Cu samples. The upper portion of the table contains results for the sputtered samples of the present study. The lower portion of the table summarizes the results reported by Pizzini, et al. for evaporated samples.IS Since Pizzini, et a]. fixed the coordination number N at 12, the data analysis for the present study treated N both as a fixed parameter and as a variable to be fit. Sample Cu Data Co Data Co A/Cu A N R (A) 02 (A'Z) N R (A) 62 (A‘z) 15/9 11.16 2.551 0.0005 0.085 1 1.61 2.513 0.0009 0.005 =12 2.551 0.0010 0.134 =12 2.512 0.0011 0.019 15/16 12.75 2.551 0.0009 0.213 12.39 2.512 0.0013 0.029 =12 2.551 0.0005 0.253 =12 2.512 0.0011 0.042 15/25 10.90 2.553 0.0000 0.167 N/A N/A N/A N/A =12 2.553 0.0006 0.258 N/A N/A N/A N/A 15/35 1 1.58 2.557 0.0003 0.035 11.69 2.512 0.0007 0.006 =12 2.557 0.0006 0.048 =12 2.512 0.0009 0.015 Pizzini, et a1.15 in-plane =12 2.54 |02|< N/A =12 2.52 0.0003 N/A 0.0005 to 0.0023 out-of-plane =12 2.55 0.0000 N/A =12 2.50 0.0008 N/A to to 0.0005 , 0.0023 79 Since the samples were mounted at 45° to the beam, the differences that Pizzini, et al.15 found for in-plane versus out-of-plane nearest neighbor distances R cannot be confirmed. However, this 45° mounting makes their cosine squared polarization factor equal 0.5, and thus agreement exists if the distances determined here are the simple average of their distances. Here Co R = 2.51A and is the average of the in-plane R = 2.52A and out-of-plane 2.50A values found by Pizzini, et al. The Cu R values found in this experiment appear to show a pattern, but appropriate error limits for the R values must be established before any interpretation. These errors are believed to be better than the often quoted generic i0.02A based upon several factors. First, increasing the momentum range used in analyzing Cu data by nearly 25% on several samples changed the distances by a maximum of 0.002A. Second, with N = 12 and 02 fixed at its optimum value, a systematic variation of R caused the goodness of fit parameter < x2 > to increase an order of magnitude for a change of i0.006A. Consequently, the NN distance results are good to better than i0.01A, and the reader is permitted to do the necessary rounding of values. Returning to the pattern of R values for Cu, R for the thickest Cu layer is essentially the 2.556A of bulk Cu while the two thinner Cu layers have a value nearer the average of this bulk value and the 2.54A in-plane value found by Pizzini, et al.15 for samples of comparable thickness. Thus, in the thickness regime where the samples of the present study and the evaporated samples of Pizzini, et al. are comparable, the observed R values are consistent. Outside this thickness regime, R for the thicker Cu layers approaches that of pure Cu for the multilayers studied here. The results for disorder in the layers are also consistent with the Pizzini, et al.‘5 80 finding that Co layers are more disordered than Cu layers. First, there is the qualitative finding about the data momentum ranges noted earlier. The C0 oscillations typically become noisy above lOA'l while Cu oscillations continue to 11.5A'l or greater. Both the Cu and Co standards give well-defined oscillations to nearly 14A’1. Second, there is quantitative agreement in the 02 values. Whether the number of nearest neighbors is varied or fixed at 12, 0'2 values obtained are similar to theirs. 6.3.3 XRD and AXRD Results XRD data were obtained using two different sources. Initial confirmation of layering was obtained from low angle XRD data taken with the Rigaku system. AXRD data were obtained at NSLS beamline X6B using two wavelengths. The wavelengths were calibrated using absorption of standard foils and their reproducibilities were tested by determining the location of the <400> line from the Si substrates. The reproducibility of the Co edge setting was 0.03% or better for the present data. AXRD data provide a valuable check upon XAFS results that indicate altered R values. All the Co/Cur samples show an average dominant Bragg line that is consistent with a fee structure having <111> fibre texture and a lattice spacing similar to Cu. This line's location should be independent of the in-plane changes in R and should vary systematically with the differences in Co layer thickness relative to Cu layer thickness. Denoting the location of this dominant line in terms of an atomic layer distance produces a quantity independent of x-ray wavelength. The results from both the present study and that of Pizzini, et al.15 are consistent with a linear dependence of upon the relative amount of Cu in the bilayer unit though the slopes from the two studies differ. 81 Despite this difference, the basic agreement for the two sets of Cu/Co samples is good. The agreement for F WHM values of the average Bragg line from the two studies is not as good. Pizzini, et al. find values of 0.l9°-0.23° while those of the present study for the more comparable x-ray wavelength have values of 0.25°-0.36°. However, even 0.36° gives a coherence length over 10 bilayer units long which suggests that the samples are of good quality. Table 6-3 summarizes the results associated with line positions found in the AXRD study. I<> denotes the average Bragg line while I‘ and 1+ denote satellites occurring at smaller and larger 29 values, respectively. The calculated values given in the table were obtained using a simple step function model for the interfaces. Deviations from this simple model typically have pronounced effects upon line intensities and only minor effects upon line locations.l7 Values of the anomalous scattering corrections used in the calculation were obtained from the f-prime routine in the General Structure Analysis System (GSAS) program.‘8 Bulk parameters were used for planar spacings normal to the plane of the layers, and a change in <111> planar atomic density was used to obtain the values for in—plane distortions consistent with the XAF S results of Pizzini, et al.15 The effects of variations in resetting the x-ray beam energy were determined by doing the model calculations for a range of energies. Neither the planar density changes nor the 0.03% variation in resetting the beam energy thus changing the wavelength produced significant changes in the calculated locations of the lines. Experimental 29 values marked with an * have unusually large intensities and may be associated with the presence of more than a single satellite line. Experimental values for the bilayer distance 82 Table 6-3: Summary of d-spacings from XRD and AXRD data. 10 denotes the dominant line; 1' denotes the satellite located just below the dominant line while I+ denotes the satellite just above it. A is the measured value of the bilayer spacing. The * indicates unusually large intensities and the ** indicates the sample with a 60A Fe buffer and a 50A Cu cap. (The buffer is the first layer sputtered; the cap is the last). The procedure for calculating positions is described in the text. @ 7.0 keV @ 7.689 keV Co A/Cu A Rigaku Calculated Exp. Calculated Exp. data 29 29 29 29 15/9 I 46.30 46.28 41.91 41.88 10 50.97 50.72 46.12 45.89 F 55.72 55.08 50.51 --- extra lines 59.7? 43.33,53.63 25.1 A (A) 25.5305 15/9.5** I 46.40 46.12 42.02 41.74 10 50.96 50.77 46.12 45.90 1+ 55.88 --- 50.37 49.72* extra lines 47.85? 23.8 A (A) 26.33213 15/16 I 47.18 47.08 42.77 42.68 10 50.82 50.62 45.98 45.80 I+ 54.52 54.19 49.38 -- ‘ extra lines 31.9 A (A) 31.6:03 4/25 I 46.70 --- 42.30 42.16 10 50.38 50.32 45.60 45.53 1+ 54.41 5505* 49.20 4969* extra lines 58.84 53.14 29.1 A (A) 83 A are also given in Table 6-3. Low angle XRD data using the Rigaku powder diffractometer produced only two lines; thus, no error estimates are available. Experimental values for A from the higher angle data, with an error estimate, are given when three line locations are available. Results in Table 6-3 for the three 15A Co samples at the lower energy form a regular pattern. Experimental 29 results are systematically smaller than predicted with 29 for 10 being about 02° below its predicted value. This difference is within the combined error associated with beam wavelength error and sample alignment error. For the two samples having three lines, the A values determined from higher and low angle data agree within experimental error. Moreover, these experimental values differ from the nominal values by less than one monolayer. Suggestive features of this pattern also appear in the data at higher energy. The lack of 1+ lines except for the one suspect apparent I+ line, suspect based upon intensity considerations as indicated by the *, prevents full completion of the pattern. AXRD results for the 4A Co sample are quite different. Only one satellite line is observed unambiguously, I" at the higher energy. Both the strong intensity of an apparent I+ line, and the fact that it is the only line in Table 6-3 whose experimental location is greater than its predicted value make its identification suspect. A strong extra line is observed at both energies. Converting the locations of the apparent 1+ line and the extra line into d-spacings reveals that the former is within 0.01A of the value for the bulk hcp <101> Co line while the latter is within 0.01A of the value for a bulk fcc <200> Cu line. At both energies the d-spacing associated with the 10 line is within 0.005A of the 84 value for the bulk Cu <111> line. These higher angle results strongly suggest that this sample might be more correctly described as being a mixture of fee Cu and hcp Co. Any attempt to upgrade this suggestion of a Cu and Co mixture to a conclusion must address several conflicting considerations. From a practical perspective, a 2 ML Co layer must be all interface, and from a fundamental perspective there are not enough Co ML's to establish a fcc structure. Starting with the latter point and arguing that the Co layer has a basic hcp structure with Cu impurities might support the appearance of a hop line. However, actual separation of this structure from Cu, even with the addition of thickness fluctuations that increase local sizes to several times the 2ML's, would be associated with a serious line broadening that would render the line unobservable. Moreover, this structural model ignores the fact that low angle data confirm some form of layering. Since the small fraction of Co ML's in the bilayer unit would cause the Cu <111> line and a composite <1 1 l> line to be nearly identical, using the position of this single line to distinguish between pure Cu and a multilayer is not reliable. Identifying the observed <111> Bragg line as such a composite line leads to structural coherence over 7 or more bilayers thus eliminating any line broadening concerns. A strong <200> Bragg line is also observed for the 4A Co sample. This line's presence indicates the sample's <1 1 1> fibre texture is weaker and raises another possible source for the unusual strength of the apparent 1+ line. A layered structure would produce satellites on a composite <200> line. Given the nominal bilayer distance of 29A, the 1‘ satellite on the <200> line would be located within 0.2 degrees of the I+ satellite on the <111> line. Moreover, a general pattern of satellite intensities discussed in the following 85 paragraph suggests that falsely identifying a <200> I“ satellite as a <111> I+ satellite would lead to an increased intensity at the higher beam energy. Such an increase is observed. Following is a reprisal of the significance of line locations for structural characterization of the XAFS samples. First, layering is confirmed and bilayer distances determined from low and higher angle data agree within experimental error. One datum questioning this simple result is related to an unusually strong line at 49.72° for the buffered/capped sample at higher beam energy. However, given the empirical presence of a strong line at 49.69° for the Co 4A/Cu 25A sample at this same beam energy and the usually strong intensity of this questionable datum, a simple I+ identification is suspect. Second, the Co ISA/Cu 9A sample has a weak extra line located near the position expected for a <200> line. The stronger <200> line observed for the Co 4A/Cu 25A sample provides a "standard" for this weak extra line. It is unclear whether the observed shift from the "standard" location is associated with a composite Cu and Co <200> line or simply the difficulty in correctly locating such a weak line. In either case, the very weakness of this line in the XAF S sample is direct confirmation of a strong <111> fibre texture that typically accompanies good layered structure in sputtered Co/Cu samples. In principle, a comparison of the satellite intensities obtained from the simple model calculation with the experimental intensities could provide an additional structural test. The calculated intensities can be viewed as an upper limit for actual results since deviations from an abrupt step interface typically decrease satellite intensities. Using the distortions suggested by Pizzini, et al.” produce calculated intensity changes of less than i20% relative to the undistorted samples. A double normalization of the satellite 86 intensities, first to the beam intensity to remove beam-decay effects and second to the I<> intensity to remove any orientational texture effects leads to the following remarks. First, the simple model correctly predicts general intensity trends. It predicts that I' intensities increase by a factor of 2.3 to 2.9 as the beam energy is increased from 7 to 7.69 keV. The observed increase factors are between 1.7 and 2.1. The model predicts I+ intensities will be less than those of I“ at the lower energy and will decrease by an additional factor of 2.5 or more at the higher energy. The I+ intensities are smaller at the lower energy and no l+ lines are unambiguously observed at the higher energy. Second, all observed satellite intensities are near the upper limit values predicted by the simple step model. This result is consistent with the inferences of sharp interfaces which arise from NMR studieswm Third, lines marked by an * in Table 6-3 have intensities that are at least 300% greater than the values predicted by the simple model. Such large intensity discrepancies contrast sharply with the general consistency of calculated and experimental results and firrther support the alternate identification of these lines. 6.4 Co/Mo Co/Mo multilayers provide an opportunity to study the effects of spacer layers which have a completely different structure from the Co layers. Not surprisingly, this system exhibits significantly different magnetic behavior than the other two Co/X systems. Unfortunately, comparison of two published studies of sputtered Co/Mo multilayers reveals limited agreement and many significant differences in both magnetic behavior and structure. Both studies find highly disordered layers and a near loss of ferromagretism when the Co layer is about 10A or less in thickness. Sato attributes this loss to a "dead 87 layer", i.e., Co atoms near the Mo layers have no moment or a reduced moment.19 Wang, et al. report the presence of a non-magnetic compound——e-Co7Mo(5.20 For thicker Co layers, these studies report different structures and different magnetic behavior; however, it should be noted that Sato fabricated his samples using magnetron sputtering while Wang, et al. used focused ion sputtering. In each study, XRD produces low angle peaks that are consistent with a layered structure. For bilayer thicknesses less than 40A, Sato'sl9 higher angle XRD data show a single dominant peak whose location is dependent upon the relative portions of Mo and Co in the bilayer unit. For multilayers having equal amounts of Mo and Co, this peak location is constant. For samples with A = 60A, this single peak has split into Bragg peaks characteristic of the Mo and Co forming the bilayer. Wang, et a1.20 find isolated Mo<110> and Co<111> peaks for all of the samples that were not disordered. In addition, a line from e-Co7Mo6 is observed for many samples, and the <002> peak for hcp Co is observed for thicker Co layers. Their TED data reveal lines consistent with bcc Mo and 3 lines associated with e-Co7Mo6. The dependence upon layer thickness for the room temperature saturation magnetizations of the two studies is also different. For samples with equal amounts of Co and M0 in the bilayer, the studies agree that 10A Co layers have a saturation magnetization less than 10% that of bulk Co; however, Sato19 reports a slow rise with increasing Co layer thickness, still less than 40% of bulk for 30A Co layers, while Wang, et al.20 report a rapid rise that reaches > 80% for 20A Co layers and 2 95% for 30A Co layers. The results for samples with differing amounts of Co and M0 in the bilayer vary somewhat in details but produce similar differences. 88 In the remainder of this section, the results of XRD, TED, XAF S, and magnetic studies of CofMo multilayers fabricated for the present study will be presented and compared to the results from the other previously cited studies. 6.4.1 Samples Fabricated The Co/Mo multilayers were fabricated using the previously described sputtering system. Sapphire and cleaved NaCl were used as substrates. The number of Co ML varied from 2-28, and the number of Mo ML, from 5-14. The same bilayer unit was made for samples on both substrates, but the total number of repeated cycles differed resulting in different total thicknesses. Samples on sapphire had a total thickness of about 2000A while that for samples on NaCl substrates was about 500A. Samples having the same bilayer dimensions were prepared consecutively. Individual layer thicknesses were chosen to match an integral number of monolayers (ML) for the metals: 2.25A and 2.05A were used as the nominal ML thickness for Mo and Co, respectively. 6.4.2 Magnetic Studies Portions of the sapphire substrate samples were used for magnetization measurements in a DC SQUID magnetometer. The field was typically applied parallel to the multilayer film. The measurements were performed at 5K to determine if the loss of magnetism previously mentioned resulted from a lowering of the Curie temperature. The magnetization results for Co/Mo samples with equal layer thicknesses from a number of studies is contained in Figure 6-1 while Figure 6-2 shows the field dependence of the magnetization for two particular samples in the current study. All values have been normalized to the saturation magnetization at 5K for pure Co (1450 emu/cm3). The bulk Co sat Mm/M 1.0 0.9 - 0.8 0.7 0.6 0.5 0.4 , 0.3 > 0.2 0.1 89 - 0.. .u . Sato 5K Sato Rm Tmp Wang Rm Tmp 0 0 d -—+— Present study 5K ,,1 i . . , A, . . ;.___.L— 5'6'7‘8’9 101112131415 Figure 6-1: Normalized saturation magnetization vs. nominal Co layer thickness for Co/Mo multilayers with tMo = tcO. In addition to data from the present study, the data of Wang, et al.20 and Sato19 are presented. points connected by broken lines are the room temperature saturation magnetization data from Wang, et al.20 and Sato.19 The single triangle datum is 5K data from Sato. The solid circles are data at 5K for the Co/Mo samples of the present study. The values of Wang, et al. approach pure bulk values at room temperature while the values in this study remain substantially below that limit at 5K where the largest values occur. Comparison with Sato is complicated by the different temperatures for the measurements. The data from this study have a general dependence upon layer thickness comparable to that reported by Sato. Since the values would be expected to decrease at higher temperature, the agreement should improve. However, Sato's single datum at 5K is about 40% smaller than the corresponding result in the present study; therefore, a believable comparison requires actual data at similar temperatures. 90 0.8 ' I Y I I I r I T I bulk Co sat M/M 60 H (kOe) b) o L O < :4 3 _ Ea . o 10 20 30 40 50 60 H (kOe) Figure 6-2: Field dependence of the normalized magnetization for the a) 14ML/14ML and b) 2ML/8ML Co/Mo multilayers on A1203 substrates. The diamagnetic contribution of the substrate is fit with a straight line to determine the y-intercept thus providing the saturation magnetization of the multilayers. 91 In addition to these samples of equal layer thickness, several samples of unequal layer thickness were examined. The normalized saturation magnetizations for two samples with very thin Co layers are 0.26 and 0.023 for the 2ML/10ML and the 2ML/8ML samples respectively. This loss of magnetism is even greater than that for the 5ML/5ML sample. Also, the 28ML/14ML sample with a C0 layer at least twice as thick as that of any other sample studied here has a saturation magnetization equal to that of bulk Co. 6.4.3 XRD Results Low angle XRD data for the Co/Mo multilayers in the present study confirm that the samples are layered with the actual bilayer distances being within i3% of the nominal values. The locations of all XRD peaks are independent of the substrate, and the thinner samples give weaker signals as expected. As the bilayers of these Co/Mo samples vary from SML/SML to 28ML/14ML, the higher angle XRD data have a progression seen for 1 The bilayer unit progresses from being highly numerous metallic multilayers.2 disordered to being a well-defmed average crystalline unit and finally to being a crystalline unit with resolved constituent Bragg lines as seen in Figure 6-3. The SML/SML sample has a weak and broad line consistent with a d-spacing of between 2.135 to 2.156A. The 7ML/7ML sample has a well-defmed line with a d- spacing of 2.145A and four satellite lines, 2 below and 2 above. This d-spacing is consistent with a weighted composite line from the d-spacings of Co and Mo and is hereafter denoted as the dominant line . As the bilayer thickness increases, this same pattern is maintained. For samples with equal amounts of Co and Mo, the location of all 92 1 04 AVC CO Mo 5 {002} ? {110} e. 2 ~ .-'=. 5 =. o . u: E f: 103 L10MU10ML5 28ML/14ML 102» Intensity (arb. units) 101 _. 10°e , 32 34 36 38 40 42 44 46 48 50 52 54 29 (deg) Figure 6-3: Comparison of the higher-angle XRD data for the 5ML/5ML, lOML/lOML, and 28ML/14ML Co/Mo multilayers. lines is consistent with 2.139 i 0.007A, and the satellite line locations shift consistent with the different bilayer thicknesses. XRD data for the 14ML/14ML sample have this same pattern of locations although there is a slight change in the progression of intensities. For this sample, satellite lines are located very near the d—spacings for pure Co and pure Mo lines, and the presence of pure element lines might account for this slight intensity change. The higher angle XRD data for the 28ML/14ML sample are clearly different. Individual lines from Co and Mo are resolved and are flanked by satellites. The coherence lengths seem to indicate this same progression. For the samples with one Bragg peak, the coherence is quite large. As the individual layers begin to display characteristics more typical of a bulk nature, the coherence begins to decrease. Given the possibility of strain in these multilayers, the preceding d-spacings are consistent with either a fcc<111> or hcp<002> line for Co and a bcc<110> line for Mo. 93 Bragg lines having other indices are not observed in the XRD data which is consistent with the high degree of fibre texture typical of sputtered metallic multilayers. Most of the samples with bilayer distances of 7ML/7ML or greater show second order XRD effects of the line and satellites that confirm the previous interpretation. The 14ML/14ML sample has very poorly resolved second order effects, and the 28ML/14ML sample has only the two pure element lines as its second order effect. For the two samples with very thin Co layers (2ML) and relatively thick Mo layers (8ML and 10 ML), XRD data clearly indicate that the samples are crystalline including the Co layer. These samples likewise show second order effects of the line and satellites. See Table 6—4 for a summary of the XRD results. Table 6-4: XRD structural results for Co/Mo multilayers. denotes a single line interpreted as a composite line composed of the Mo<110> line and a C0 line. The relative error for these d-spacings is $0.003 A. The * denotes a broad, weak line which is more likely an indication of the nearest neighbor distances in an amorphous sample. Co/Mo A (A) d-spacing of # peaks from bilayer Coherence (in ML) Nom. Meas. (A) satellites harmonics Length (A) 2/8 22.1 21.8 2.169 1 3 130 2/10 26.6 25.9 2.180 2 2 230 5/5 21.4 21.8 2.135 to 2156* 0 0 15? 7/5 25.6 25.8 2.126 3 3 290 7/7 30.2 30.1 2.145 4 3 310 10/10 43.0 42.0 2.135 4 2 260 14/14 60.2 61.4 2.132 4 3 180 28/14 88.9 87.5 Mo<110> 2.208 2 4 100 C0 2.045 1 140 or Co<002> 94 6.4.4 TED Results The TED studies were performed with the VG H8501 FESTEM previously described. The TED data are consistent with the structural progression suggested by the XRD data. All the TED patterns are dominated by rings. Figure 6-4 shows that for the SML/SML sample, the pattern consists of one intense but broad ring and two weak, diffuse rings indicative of an amorphous structure. The 7ML/5ML sample reveals some interesting features. A TEM image of the 7ML/5ML sample shown in Figure 6-5 and the two diffraction patterns given in Figure 6-6 indicate a mixture of crystalline and amorphous structures. Figure 6-7 shows that the TED patterns yield a sequence of strong Bragg lines that are consistent with a bee structure for all samples with a bilayer distance of 7ML/7ML or greater. Typically 56 lines in the sequence are evident and attributed to the Mo layers. TED patterns for the samples with thin Co layers (2ML/8ML and 2ML/10ML) indicate that these samples are crystalline with 5 and 6 Bragg lines observed respectively that are consistent with a bcc structure. These results are again consistent with the XRD data. Also, none of the TED data exhibit non—bcc lines so TED provides no direct evidence for the presence of Co-Mo compounds. 6.4.5 XAFS Results XAF S data for the Co K edge were collected at the NSLS on beam line X23A2 using a Lytle detector operated in fluorescence mode. The samples were mounted at 45° to the beam and a ux = 3 Mn filter was used. The detector which monitored the incoming beam (10) was filled with a mixture of flowing N2 and He, and the Lytle detector was filled with flowing Ar. A harmonic check was performed as described in section 6.3.2. Typically 5 : t - .‘1"n..-.‘.lh£u- 25:51 95 Figure 6—4: TED line scan for Co/Mo (5 ML/SML). Figure 6-5: TEM image of Co/Mo (7 ML/SML). The nominal magnification is 50 K. Figure 6-6: TED line scans from different regions of the same Co/Mo (7ML/5ML) sample illustrating the varying degree of crystallinity. 21) b) 97 a) jIII \k b) I I Figure 6-7: a) TED line scan for Co/Mo (7ML/7ML) and b) TED line scan for Co/Mo (10ML/10ML). 98 5 I I l I T l 1 I l l I Tf 4” M ‘ 3t _ :3 2 >4 ,_J O _ I I 1 i l 1 1 1 l 1 l 1 l i 1 A 1 r l I 1 4 1 A T 7.5 7.6 7.7 7.8 7.9 8.0 8.1 8.2 8.3 8.4 8.5 8.6 8.7 Energy (keV) Figure 6-8: Combined raw data from 5 fluorescence scans of the 7ML/7ML multilayer. scans were done with 2-4 more scans done for those samples with thin Co layers. Results from different scans were combined during the analysis using MacXAFS v. 3.6. Data from a sputtered 2000A Co film was used as the standard in the analysis. Sample raw data for the 7ML/7ML sample is given in Figure 6-8. Qualitatively these data reinforce the validity of the structure development with increasing bilayer thickness as described in section 6.4.3. The XAFS data for the SML/SML sample consist of a single decaying oscillation while data for the 7ML/5ML sample are consistent with a mixture of amorphous and crystalline Co. Data for the 7ML/7ML sample and those samples with thicker Co layers have oscillations characteristic of hcp or fcc Co or a mixture of the two. These data indicate the local structure as well as the average long range structure changes from amorphous to crystalline as the bilayer thickness is increased from 5ML/5ML to 7ML/7ML in Co/Mo. 99 0.06 0.05 e 0.04 — 0.03 » 0.02 0.01 0.00 ~ ~ P , ~ ~. .r‘----"’4“ I ' “ -0.01 -0.02 1" ; 003 ~ ' 7 Tn -004 »— -0.05 CHI 3 4 5 6 7 8 9 10 ll 12 l3 l4 {Fr-'1 " ' b) 12 11 10~ RSP OHNUJ-hLIIO‘QOOC Figure 6-9: Comparison of the CHI and RSP data for the 5ML/5ML and 7ML/7ML samples clearly shows the transition from an amorphous to a crystalline structure. 100 1800 ' I T 7 I I ’ l I I ‘ 12 . D . 1600 - ' I r 11 ’ ~ 10 1400 - . a? . I 9 E 1200 h j 8 o I I . 3 1000 — I _. 7 8 ~ 2 v 800 r I 1 6 ‘5 . E 600 ~ I If M5,, 7 5 . . 4 4 400 h . o N j 3 ., 200 — _‘ E 0 1* 1 , 1 r l r l 1 l 1 3', 0 5 10 15 20 25 30 ' tCo (ML) Figure 6-10: Saturation magnetization and number of nearest neighbors as a function of 0 nominal Co layer thickness. The solid data points are for samples of equal layer thickness. The unfilled data points are for the 28ML/14ML and 2ML/8ML multilayers. A direct comparison of the CHI data and its transform for the 5ML/5ML and the 7ML/7ML multilayers in Figure 6-9 clearly denotes this transition. A correlation between structure and magnetization is shown in Figure 6-10. Both the coordination number N as measured relative to the Co standard and the saturation magnetization M5,, are shown as functions of Co thickness. The scaling of N and M53, are very similar indicating a clear correlation between the local structure and the loss of magnetism. Ignoring the 2ML/8ML multilayer for now, it is clear that the development of hcp or fcc Co coincides with the increase in saturation magnetization. Figure 6—11 shows the Fourier transforms, i.e., the pseudo-RDF's, of all the samples included in the XAFS study. It clearly reveals the different structure of the 2ML/8ML sample. The other structural studies as well as the XAFS data indicate that the Co layers are crystalline in this sample even though it has the greatest loss of magnetism. b) Figure 6-11: Comparison of the CHI and RSP data for the Co standard and all of the CHI RSP 0.10 0.08 ~ 0.06 0.04 ~ 0.02 ‘ 0.00 '. -0.02 -0.04 -0.06 -0.08 101 —- Co STD ‘ --------- 28 ML/14 ML ‘ —- 10 ML/lO ML; --------- 7 ML/7 ML —— 7 ML/S ML --------- 5 ML/S ML 2 ML/8 ML ‘3‘4'5 6‘7'8‘9‘10‘1'11‘2V‘l‘3'1'4 — Co STD ------ 28 ML/l 4 ML — 10 ML/l 0 ML ------- 7 ML/7 ML ——- 7 ML/S ML ------- 5 ML/S ML 2 ML/8 ML Co/Mo multilayers included in the XAFS study. 102 Unfortunately, the structure of these Co layers is not known. FEFF models of pure hcp, fcc, and bcc Co do not produce results that even vaguely resemble the data. Another FEFF model substituting Mo atoms for Co atoms in an hep-Co structure also failed to match the data. However, attempting to model data for only one sample is an extremely difficult task at best. Additional data for samples with x ML Co/y ML Mo with x = 2-5 and y 2 7 are needed to establish experimental trends so that FEFF may be effectively used to determine the structure of Co in thin layers. lC.L. Foiles and J .M. Slaughter, J. Appl. Phys. 63, 3209 (1988). 2s.s.P. Parkin, et a1., Phys. Rev. Lett. 64, 2304 (1990). 3s. Pizzini, et al., Phys. Rev. B 46, 1253 (1992). 4S.F. Cheng, et al., Phys Rev. B 47, 206 (1993). 5CL. Foiles, R. Loloee, and T.I. Morrison, Mat. Res. Soc. Symp. Proc. 307, 149 (1993). 6RA. Schroeder, private communication with the authors listed in reference 5. 7MA. Howson, 13.1. Hickey, J. Xu, and D. Greig, J. Magn. Magn. Mater. 126, 416 (1993) 8D.H. Mosca, F. Petroff, A. Fert, P.A. Schroeder, W.P. Pratt and R. Loloee, J. Magn. Magn. Mater. 94, 1 (1991). 9F.J. Lamelas, C.H. Lee, Hui He, W. Vavra, and Roy Clarke, Phys. Rev. B 40, 5837 (1989) 10K. Le Dang, P. Veillet, Hui He, F .J . Lamelas, C.H. Lee, and Roy Clarke, Phys. Rev. B 41, 12902 (1991); K. Le Dang, P. Veillet, P. Beauvillian, C. Chappert, Hui He, F.J. Lamelas, C.H. Lee, and Roy Clarke, Phys. Rev. B 43, 13228 (1991). llH.A.M. de Gronckel, K. Kopinga, W.J.M. de Jonge, P. Panissod, J.P. Schille, and F.J.A. den Broeder, Phys. Rev. B 44, 9100 (1991). 12C. Meny, P. Panissod and R. Loloee, Phys. Rev. B 45, 12269 (1992). l3T. Valet, P. Galtier, J.C. Jacquet, C. Meny, and P. Panissod, J. Magn. Magn. Mater. 103 121, 402 (1993); C. Meny, P. Panissod, P. Humbert, J .P. Nozieres, V.S. Speriosu, B.A. Gurney and R. Zehringer, ibid., 406 (1993). 14P. Panissod and C. Meny, J. Magn. Magn. Mater. 126, 16 (1993); Y. Saito, K. Inomata, A. Goto, and H. Yasuoka, ibid, 466 (1993); M. Suzuki, Y. Taga, A. Goto, and H. Yasuoka, ibid, 495 (1993). 15 S. Pizzini, F. Baudelet, A. Fontaine, M. Galtier, D. Renard, and C. Marliere, Phys. Rev. B 47, 8754 (1993). 16CL. Foiles, R. Loloee, and T.I. Morrison, Mat. Res. Soc. Symp. Proc. 307, 149 (1993). RF“- l7C.L. Foiles, private communication. ”Ac. Larson and RB. Von Dreele, Los Alamos Natl. Lab. LAUR86-748. l9Nohoru Sato, J. Appl. Phys. 63, 3476 (1988). 20Y. Wang, F.Z. Cui, w.z. Li and Y.D. Fan, J. Magn. Magn. Mater. 102, 121 (1991). ”BY. Jin and 1.13. Ketterson, Adv. in Phys. 38, 189 (1989). CHAPTER 7 FE/SI AND F E/ {FESI} MULTILAYERS 7.1 Introduction During the 1980's and early 1990's, Fe-Si multilayers (Fe/Si) with thick Si layers, typically 35A or greater, were studied extensively and seemed to be well understood. In one study using Mossbauer spectroscopy, an amorphous alloy was found separating crystalline Fe and amorphous Si (a—Si) layers. This study also determined that Fe layers less than 20A were not crystalline.l Polarized neutron-scattering studies indicated results consistent with an amorphous alloy interface having a 16A thickness. Cross-sectional TEM showed that these same samples also had well-defined and continuous layers.2 Another study determined that the Fe layers were not coupled and that approximately 8- 10A of each Fe layer was nonmagnetic.3 Later studies further evolved the picture by indicating that the Fe-Si interface was 18A thick with a compositional gradient consisting of Si-substituted crystalline Fe, magnetic amorphous, and non-magnetic amorphous layers.4 All of the interfacial thicknesses listed above are totals for the entire bilayer, i.e., the thickness of each of the two interfaces associated with each bilayer unit is half the listed interfacial thickness. More recent studies of F e-Si systems with thin Si layers, however, have indicated that the above picture is incomplete. Toscano, et al. discovered an oscillatory magnetic coupling in Fe-Si trilayers while studying the spin polarization of secondary electron emission.5 They claimed that the magnetic coupling is mediated through semiconducting a-Si even though no supporting structural data was provided. Such a spacer could be a 104 105 source of the conduction electrons required in the spacer layer for an exchange type coupling. This coupling discovered by Griinberg6 in Fe/Cr/Fe trilayer sandwiches was first used to explain the coupling of magnetic layers separated by a nonmagnetic metallic spacer. It is characterized by an energy per unit area of the form: Ec = —2A12(M1-M2)—2B12(M1-M2)2 (Eqn. 7-1) where Ec is the exchange coupling energy, M1 and M2 are unit vectors along the magnetic moments of the two magnetic layers, and A12 and B12 are the respective bilinear and biquadratic exchange coupling constants. The biquadratic term7 was found necessary subsequent to the initial finding and is required only in certain cases. Further support for exchange coupling came from a series of studies performed by a group at Argonne. For sputtered F e/Si multilayers with 30A thick Fe layers and nominal Si spacer layer thicknesses (tSi) between 10 and 40A, they reported ferromagnetic (FM) behavior for ts, < 13A, antiferromagnetically (AF) coupled Fe layers for 13A < ts, <17A, and no coupling between Fe layers for ts, > 17A.8 These magnetic properties were studied using superconducting quantum interference device (SQUID) magnetometry, longitudinal Kerr rotation, conversion-electron Mossbauer spectroscopy (CEMS), and polarized neutron diffraction. Using XRD, they determined that the Fe layers were structurally coherent for ts, < 17A. Both the magnetic coupling and the structural coherence were explained by the presence of an Fe silicide inferred from the CEMS which was reported to most likely be s-FeSi. This semiconductor would be a much better source of the required conduction electrons since its band gap energy is 0.05eV compared to the 1.12eV of a-Si. The Argonne group later indicated finding AF coupling and ”.111" “I i' ‘ 106 structural coherence up to spacer layer thicknesses of 40A when the spacer was {FeSi} .9 This spacer was made by depositing approximately one monolayer (ML) of Fe followed by approximately one ML of Si and repeated to the required thickness. Independent studies found that the magnetic coupling could be modified optically and further supported the case for exchange couplingm‘” Recently, an electron transport study of sputtered Fe/ Si multilayers by Inomata, et al. revealed a negative magnetoresistance (MR) with a significant change in its temperature dependence as the Si layer thickness exceeds 15A.12 From the temperature dependence of magnetic properties, they also inferred that the spacer layers consisted solely of s-FeSi for ts, < 15A and a combination of s-FeSi and a-Si when ts, > 15A. Nonetheless, this seemingly strong support for exchange coupling is lessened by the fact that the presence of e-FeSi was never directly confirmed and by recent questions of the validity of the optical studies.l3 Moreover, a group at Lawrence Livermore National Laboratory (LLNL) shared the results of their study prior to its publication which seemed to indicate that the magnetic coupling was dominated either by dipolar coupling related to interfacial roughness or by 4 In multilayers comprised of superparamagnetism of weakly-coupled Fe particles.l magnetic layers separated by nonmagnetic metallic spacer layers, the work of Hill, et al. also indicates that dipolar coupling energies originating from interfacial roughness can explain both FM- and AF-like coupling.” Though single Fe-Si interfaces have been studies extensively, the findings are highly dependent upon grth method and conditions. '6'” Consequently, reliable extrapolation of these results to the interfaces found in sputtered multilayers is questionable. Clearly, direct structural information of these multilayers is required to determine the nature of the 107 spacer layer and possibly the mechanism of magnetic coupling between Fe layers. In this chapter, the results of magnetic studies, XRD, TED, and XAFS of both Fe/Si and Fe/{FeSi} multilayers will be presented. 7.2 Samples Fabricated Using the sputtering system described in Chapter 2, F e/Si and F e/{FeSi} multilayers were fabricated with all Fe layers being nominally 28.7A (14ML) thick. For the Fe/Si samples, 10A 3 ts, 3 30A. The Fe/{FeSi} spacer layers were made by alternately depositing nominally 2A of Si and 2A of Fe repeatedly to obtain total spacer thicknesses (ti-85,) of 12-40A. Sapphire, crystalline Si, and cleaved NaCl and KC] were used as substrates. All of the total multilayer thicknesses were around 500A except for the Fe/Si multilayers on sapphire and Si substrates which were around 2000A in total thickness. 7.3 Magnetic Studies Portions of the sapphire and Si substrate samples were used for magnetization measurements in a DC SQUID magnetometer. The magnetization as a function of the field applied both parallel and perpendicular to the multilayers was determined at 5K. Additional measurements were made using an AC SQUID magnetometer to determine remanent magnetizations (M,) for further evidence of coupling. At low fields, the magnetic field of the AC SQUID is closer to an actual value of zero thus providing more accurate determinations of M,. For all of the Fe/Si multilayers, the saturation magnetization (Mm/V) was about 70% that of bulk Fe essentially independent of the spacer thickness. Also, the large parallel fields required to achieve saturation and the field in excess of the demagnetizing 108 Table 7-1: Magnetic properties of Fe/{FeSi} multilayers as a function of spacer layer thickness. The Msat data were obtained with the DC SQUID, and the Hsat and MrlMsm data, with the AC SQUID. Cycles 0f Msat/V pure Fe Msat/V all Fe Hsat (parallel) Mr/Msat (Fez/8180A} (103 emu/cm?) (103 emu/cm3) (kOe) 3 1.8 1.5 4 1.8 1.4 0.3 0.95 5 1.7 1.3 1.2 0.93 6 1.8 1.2 1.4 0.95 7 1.8 1.2 5.0 0.81 8 1.8 1.2 3.7 0.83 10 1.7 1.0 field required for saturation in the perpendicular geometry were consistent with the results of the previously mentioned studiess’lz. Likewise, the coupling evidenced by the Fe/ {FeSi} samples was consistent with the Argonne group's fmdings.9 A summary of the magnetic properties is shown as a function of spacer thickness in Table 7-1. For the thinner spacer layers, the results are consistent with FM coupled or uncoupled Fe layers. As the spacer thickness increases, the results are consistent with AF coupling with a gradually decreasing strength. Ifit is assumed that the Fe in the {FeSi} spacer is non-magnetic such that the saturation magnetization is normalized to the volume of the nominally pure Fe layers (V pure Fe) only, then the measured value for all spacer thicknesses is approximately equal to that of bulk Fe (1.71x103 emu/cm3) within the i 5% error in these measurements resulting from the uncertainty in the bilayer distance. Normalizing the magnetization to the total volume of Fe (V all Fe) results in a saturation magnetization which varies with spacer thickness as shown in Table 7-1. The saturation magnetizations determined from the AC SQUID data are slightly larger than those obtained from the DC SQUID data. The AC SQUID data in . ' ~ .QI‘... 109 1.14 . T 1 . . . . , , I ‘ 1.13 i 1.12 a 1“ 1H HI € 1.10 g' 1.09 F' T . 1.08 i "a I - 1.07 T ; 1.06 1.05 l i 1.04 l i 1.03 - 1.02 _ . . . . . . . M/MWc Fe sat I" Fl 1 b) 1.14 a 1.12 1.10r 1.08’ 1.06’ 1.04. 1.02" 1.00. 0.98 0.96h 094* 0.92 0.90 F M/Mg;::re e H (kOe) Figure 7-1: Field dependence of the magnetization normalized to the volume of the nominally pure Fe layers for the a) {F e2ASi2A} X 4 and b) {F e2ASi2A} X 7 spacer layer samples on A1203 substrates. These AC SQUID data give slightly larger saturation magnetizations than the DC SQUID data (see Table 7-1). '17 110 Figure 7-1 shows this slight discrepancy as well as the field dependence of the magnetization for two of the multilayers. Regardless, either of these normalizations yields different results from the F e/ Si multilayers and shows that further experiments are required to determine whether or not the Fe in the {FeSi} spacer is magnetic and if it affects the magnetization reduction for the nominally pure Fe layer. 7 .4 XRD Results The x-ray data were taken using the Rigaku system described in section 3.3.2. No differences in d-spacings greater than experimental errors were observed for the same sample on different substrates, but differences in intensities were seen. 7.4.1 Low-angleXRD Low-angle harmonics provide evidence of layer formation for both Fe/Si and Fe/{FeSi}. The number of harmonics varies from 3-8 for Fe/Si and from 2-5 for Fe/{FeSi}. The bilayer distances A were determined as discussed in section 3.5.2. However, the two systems differ markedly with regards to the agreement between the measured and nominal values of A. For the Fe/Si multilayers, the measured values of A are substantially lower than the nominal values. Figure 7-2 shows this variation of A as a function of ts, for those samples on sapphire substrates—those on Si substrates behaved similarly. This result was also observed by the Argonne group18 which they interpreted as evidence of interdiffusion. Yet, for the F e/ {FeSi} multilayers, the measured A agrees fairly well with the only differences attributable to random error. IE. O Bilayer Spacing (A) L (A) 59 r 57 55* 53 51 49 47 45 y. 43 41 39 37 35 240 220 . 200 180 160a 140* 120 100 80 60 40 20s 26 28 Figure 7-2: Measured bilayer spacing as a function of nominal Si layer thickness for F e/Si multilayers. The solid line indicates nominal bilayer spacings. ‘—O—.—.——o——o—o tspacer (A) Si spacer {FeSi} spacer 35 Figure 7-3’: Coherence length vs. nominal Si layer thickness for both Fe/Si and Fe/{FeSi} multilayers on A1203 substrates. The horizontal line indicates a constant value. A. . 112 7.4.2 Higher-angle XRD A single Bragg peak consistent with the <110> line of Fe is observed for all samples. For the Fe/Si multilayers with thin Si layers, this line is narrow indicating a structural coherence greater than one bilayer. In addition, weak satellites are observed. For ts, > 14A, the Bragg peak broadens, and the coherence length becomes comparable to that of a single Fe layer. No satellites are seen for these samples as well. In contrast, the Fe/{FeSi} samples exhibit structural coherence over two or more bilayers for all spacer thicknesses studied and typically have at least one satellite. Two satellites are seen for the two thickest spacer layers. Figure 7-3 shows the coherence lengths as a function of spacer thickness for all samples. The variation in planar spacing for the <110> line as a function of spacer thickness is shown in Figure 7-4 and Figure 7-5. For comparison, the accepted d-spacing for the <110> line of bulk bcc Fe is 2.027A, and a measured value using the Rigaku diffractometer for a 450A Fe film sputtered onto a NaCl substrate is 2.025A. For the Fe/ Si samples, the d-spacing is essentially constant with the exception of the thinnest Si layers which have a much greater coherence length. Simple step models used to predict XRD intensities show that not only do the maximum intensities and peak widths change when the structural coherence is increased, but the peak positions change as well. The observed change in d-spacing for the Fe/Si multilayers and the accompanying change in coherence are consistent with step model predictions. The F e/ {FeSi} multilayers do not exhibit this change in coherence; however, the planar spacings decrease with increasing spacer thickness. As stated previously in the XRD chapter, when two difierent crystalline materials with similar d-spacings comprise Tr 113 ‘r r v v y r I Y Y Y T 2.03 7 i r 2.02 d-spacing (A) 2.01 0 A1203 substrate D A1203 substrate 2'00 (different run) I 1 .99 l L l 1 l L l 1 l i l I l I l i l L J 4 l 10 12 l4 16 18 20 22 24 26 28 3O tspacer (A) b) l 40 120 A 3 .a 100 a}. —— t31=10A : :. ,’ 1: ......... tSi = 20 A e 80 , .3 '1' ' . a! 3.3137 "' 1:; , v :2; s ‘ =-.-'=.-'. s: >3 60 4:.5'25 35 ‘- H _ ,.'---. . ' . 2: S 40 i" ‘ , :3 :5 5:. bl! ' 5 ~ 20 . a”. 0 f 1 ‘ 41 42 43 44 45 46 47 48 49 20 (deg) Figure 7-4: a) Planar spacing as a function of spacer layer thickness for the Fe/Si samples and b) a comparison of the higher-angle XRD data for the ts, = 10A and ts, = 20A multilayers. 114 2.026 ~ ~ 2.024 : T i 2.022 L «4 2.020 r 1 1 2.018 P i i 2.016 P —« d-spacing (A) 2014 L * ZIHZ * r 2.010 " :1 2008 I i 1' 1 l 1 1 I l 1 l 10 15 20 25 30 35 40 tspacer (181) b) 800 700 600* 500 {FeZASiZA} x 10 400 300 {FeZASiA} x 6 200 Intensity (arb. units) 100~ {Fe2ASiZA} x 3 ;_ i l 42 43 44 45 46 A 47 48 20 (deg) Figure 7-5: a) Planar spacing as a function of spacer layer thickness for the Fe/{FeSi} samples and b) a comparison of the higher-angle XRD data for the {Fe2ASi2A} X 3, X 6, and X 10 multilayers. 115 a bilayer unit in a multilayer then one Bragg line is observed with a d-spacing that is the weighted average of the constituent components for a certain thickness range of the individual layers. Alloying Si into Fe causes a decrease in planar spacing proportional to the Si content.‘9 The observed decrease in d-spacing with increased spacer thickness is consistent with an Fe-Si alloy constituting an increasing portion of the spacer layer. It should be noted that the entire range of d-spacings in Figure 7-4 and Figure 7-5 lies within the limits of pure Fe and the F e-Si alloy with greatest Si content which are 2.027A and 2.006A respectively. 7 .5 TED Results All measurements were made using the VG HB501 FESTEM described in Chapter 5. The number of lines seen in the F e/Si multilayers vary from 14 for those samples with the thinnest Si layers to as few as 8 for the thicker Si layers as shown in Figure 7-6. All of these lines are consistent with a bcc structure. A summary of these observed Bragg lines is presented in Table 7-2 along with the results for a 450A Fe film included for comparison. The presence of the <400> and <521> lines in many of the Fe/Si samples which are missing from the pure Fe film may indicate that alloying has taken place since substitution of Si atoms for Fe atoms changes the atomic form factor for the planes containing Si. This substitution can produce stronger <400> and <521> lines than those observed in pure Fe. The number of Bragg peaks for the Fe/ {F eSi} samples do not show the same trend as the Fe/Si samples—instead the number seems more nearly constant with random differences. Given the lack of change in the coherence for the thickness range examined, 116 6!) b) J Figure 7-6: TED line scans for Fe/Si samples with nominal Si layer thicknesses of a) 12A and b) 3021. 117 Table 7-2: TED Bragg lines observed in F e/Si samples. The lines are identified by their Miller indices. Y indicates that the line is observed, and N that it is not. 28.7 A Fe/x A Si x=0 10 12 14 16 18 23 26 30 <110> Y Y Y Y Y Y Y Y Y <200> Y Y Y Y Y Y Y Y Y <21 1> Y Y Y Y Y Y Y Y Y <220> Y Y Y Y Y Y Y Y Y <310> Y Y Y Y Y Y Y Y Y <222> Y Y Y Y Y Y Y Y Y <321> Y Y Y Y Y Y Y Y Y <400> N Y Y Y Y Y N N N <411;330> Y Y Y Y Y Y Y Y Y <420> Y Y Y Y Y Y N N N <332> Y Y Y Y N Y N N N <422> Y Y Y Y N ' N N N N <510;431> Y Y Y Y N N N N N <521> N N Y Y N N N N N this result is not surprising. All of the Fe/{FeSi} samples have 10 or more lines consistent with a bcc structure. "Most" samples have at least one additional non-bcc line. If a very weak line in only one out of the three regions examined is accepted as evidence, then "most" becomes all. Two of the samples have two non-bcc lines as shown in Figure 7-7. Neither the number of these lines nor their strengths reveal any discemible correlation with the {FeSi} layer thickness. For instance, the 16A and 40A spacer layer samples are the multilayers having 2 non-bcc lines. These lines have d-spacings of 3.08A and 2.51A if the identified bcc lines are assumed to have the planar spacings of bulk Fe. The identification of the material producing these two extra lines is difficult at best. Since it seems likely that they might indicate the presence of an Fe silicide, the listings for d-spacings of known equilibrium Fe silicides from the Powder Diffraction File were compared to the preceding values. By far the best match was for that of e-FeSi—the 118 1 Figure 7-7: TED line scan for the {Fe2ASi2A} X 10 spacer layer sample with the two non-bcc lines indicated by the arrows. Tl same silicide cited in the studies mentioned previously. The <110> and <111> lines of s- FeSi have respective planar spacings of 3.16A and 2.59A. In addition, the next several lines which have significantly greater intensities in XRD powder patterns have d-spacings such that they would be hidden by the observed bcc lines. Consequently, these findings are consistent with the presence of s-FeSi although the non-equilibrium conditions associated with sputtering could result in the formation of non-equilibrium compounds for which no comparison can presently be made. TEM images show some unusual aspects. Normally, the sputtered multilayers investigated in this study appear to be quite homogeneous when observed with the FESTEM. This result is also the case for all of the Fe/Si samples and for the Fe/{FeSi} samples with tires, 3 24A. The remaining samples, however, have BF images consisting of mostly dark regions with small inclusions of significantly lighter regions. The dark 119 regions are crystalline and have the characteristics noted above. The lighter regions are amorphous and probably consist of mostly Si though this was never confirmed. Unfortunately, the x-ray detector was not working during these measurements, so EDX (energy dispersive x-ray analysis) of the composition could not be used. Later, when EDX was again available, the samples on salt substrates were observed to have suffered significant oxidation even though kept in a vacuum dessicator. Since positive identification of these regions would not provide any more information regarding the magnetic coupling, additional samples were not made. 7.6 XAFS Results A limited analysis of the XAFS data reveals several qualitative features. For Fe/Si multilayers, the EXAF S region has oscillations quite similar to that of pure Fe indicating a bcc structure but with a reduced number of Fe first and second shell neighbors which far exceeds that expected for sharp Fe-Si interfaces. In the XANES region comprised of about 40 eV above the edge, some Fe/Si samples with thicker spacer layers are missing a broad hump seen for pure Fe. For the Fe/{FeSi} multilayers, the EXAF S data appears to be the same as that for pure Fe with a decreased amplitude indicating a reduction of neighbors in the first two shells of about one-half. A variety of structural models have been used to generate theoretical XAF S data using FEFF. So far, none of these models—including sharply interfaced multilayers, Fe- Si alloys of varying composition, Fe silicides, and plausible weighted combinations of these models—have produced the observed XAFS data. '1 120 1C. Dufour, A. Bruson, B. George, G. Marchal, and Ph. Mangin, J. Phys. (Paris) Colloq. 49, C8-1781 (1988). 2C. Dufour, A. Bruson, B. George, G. Marchal, Ph. Mangin, C. Vettier, J .J . Rhyne, and R.W. Erwin, Solid State Commun. 69, 963 (1989). 3C.L. Foiles and J.M. Slaughter, J. Appl. Phys. 63, 3209 (1988). 4C. Dufour, A. Bruson, G. Marchal, B. George, and Ph. Mangin, J. Magn. Magn. Mater. 93, 545 (1991). 5S. Toscano, B. Briner, H. Hopster, and M. Landolt, J. Magn. Magn. Mater. 114, L6 (1992) 6P. Griinberg, R. Schreiber, Y. Pang, M.B. Brodsky, and H. Sowers, Phys. Rev. Lett. 57, 2442 (1986). 7M. Ruhrig, R. Schafer, A. Hubert, R. Mosler, J .A. Wolf, S. Demokritov, and P. Grl'inberg, Phys. Stat. Sol. (a) 125, 635 (1991). 8Eric E. Fullerton, J .E. Mattson, S.R. Lee, C.H. Sowers, Y.Y. Huang, G. Felcher, S.D. Bader, and RT. Parker, J. Appl. Phys. 73 (10), 6335 (1993). 91E. Mattson, Eric E. Fullerton, Sudha Kumar, S.R. Lee, C.H. Sowers, M. Grimsditch, S.D. Bader, and RT. Parker, J. Appl. Phys. 75, 6169 (1994). loJ.E. Mattson, Sudha Kumar, Eric E. Fullerton, S.R. Lee, C.H. Sowers, M. Grimsditch, S.D. Bader, and F .T. Parker, Phys. Rev. Lett. 71, 185 (1993). “B. Briner and M. Landolt, z. Phys. B92, 137 (1993). 12K. Inomata, K. Yusu, and Y. Saito, Phys. Rev. Lett. 74, 1863 (1995). 13Eric E. Fullerton, private communication. l4Alison Chaiken, private communication. 15E.W. Hill, S.L. Tomlinson, J .P. Li, J. Appl. Phys. 73 (10), 5978 (1993). l6J. Alvarez, A.L. Vazquez de Parga, J.J. Hinarejos, J. de la Figurera, E.G. Michel, C. Ocal, and R. Miranda, Phys. Rev. B 47, 16048 (1993); J. Vac. Sci. Technol. A 11, 929 (1993). 17Papers in Silicides, Germanides, and Their Interfaces, Mater. Res. Soc. Proc. 320, (1994) Il— 121 18Eric E. Fullerton, J.E. Mattson, S.R. Lee, C.H. Sowers, Y.Y. Huang, G. Felcher, S.D. Bader, and F .T. Parker, J. Magn. Magn. Mater. 117, L301 (1992). 19W.B. Pearson, Handbook of Lattice Spacings and Structure of Metals and Alloys, Vol. 2, Pergamon Press (1967). CHAPTER 8 SUMMARY AND CONCLUSIONS 8.1 Co/Cu Multilayers XAFS data for sputtered Cu/Co multilayers give nearest neighbor distances and measures of layer disorder which are consistent with the earlier results of Pizzini, et al. for evaporated multilayers. The use of anomalous dispersion to enhance the x-ray contrast between Cu and Co supplement the XAF S data in two respects. First, the actual bilayer distances of the sputtered samples have been measured and confirm the use of nominal bilayer spacings in describing experimental trends in behavior. Second, both the Bragg lines and their general intensity patterns suggest that the XAF S data are free of any complications associated with significant deviations from either sharp interfaces or strong <111> fibre texture. 8.2 Co/Mo Multilayers The saturation magnetization shows a systematic decrease from bulk value for thick Co layers to a nearly complete loss of magnetism for thin Co layers. This loss occurs for both crystalline as well as disordered samples. These results conflict with those of Wang, et al., but are generally consistent with those of Sato. Using XRD and TED, a structural progression is observed for the multilayers. For equal layer samples, a transition from amorphous, to crystalline with a dominant composite Bragg line , and finally to Bragg lines for the individual elements is seen as the bilayer thickness increases. XAF S data document that this transition from disorder to crystalline order occurs rapidly as layer 122 ,. J 123 thickness increases from SML to 7ML. TED data provide no evidence for the presence of Co-Mo compounds. These structural findings are consistent with those of Satol for Co/Mo prepared by dc magnetron sputtering and are completely inconsistent with the structural features reported by Wang, et al.2 for Co/Mo prepared by focused ion sputtering. These observations seem to indicate that magnetron sputtering and focused ion sputtering form Co/Mo multilayers with different structures. XAFS data show a clear correlation between local structure and magnetic behavior. For equal layer samples, XAF S data clearly show that the development of crystalline Co into an hcp or fee structure or a mixture of these coincides with an increase in saturation magnetization. For the one crystalline sample (2ML/8ML) with thin Co layers, XAF S data reveal that the Co layers have a different structure from that of the other Co/Mo multilayers. Though several plausible structures for this sample were modeled using FEFF, none of the modeled results were consistent with experiment. 8.3 Fe/Si and Fe/{FeSi} Multilayers XRD data establish a consistent pattern of structural coherence in Fe/Si and Fe/{FeSi} multilayers. This coherence occurs for a limited range of nominal Si layer thicknesses (ts, S 15A) and for a much larger range of {FeSi} thicknesses “{FeSi} 2 40A). XRD data provide no direct evidence of F e—silicides, but TED data exhibit non-bcc lines that are such evidence. These lines occur in Fe/{FeSi} multilayers, do not occur in Fe/Si multilayers, and give a plausible match to the Bragg lines of one equilibrium silicide. TEM images of the F e/ {F eSi} multilayers indicate that a two phase material is formed for thicker {FeSi} spacer layers. 124 These results suggest that inferring the presence of silicides from detailed features in XRD data or macroscopic properties has limited value. For samples having structural coherence over more than one bilayer, the coherence lengths determined from XRD data are comparable for Fe/Si and Fe/{FeSi} multilayers but Fe-silicide lines are only observed in the latter samples. Systematic changes in the position of the <110> Bragg line are consistent with structural coherence but give no evidence for either silicides or a change at the {FeSi} spacer layer thicknesses for which TEM images show evidence of two phases with different structures. XAFS data indicate a reduction in the Fe coordination numbers for either Si or {F eSi} spacer layers. The results of saturation magnetization measurements indicate that the Fe in the {FeSi} spacer layers exhibits different magnetic behavior from that of the nominally pure Fe layers. The underlying cause of these differences has not yet been identified. 8.4 Future Investigations 8.4.1 Co/X Multilayers With the exception of Co/Mo, the studies of Co/X were largely unsuccessful in determining any simple correlations between structure and magnetic behavior. In the case of Co/Mo, more samples with thin Co layers (2ML—5ML) are needed to establish experimental trends. FEFF modeling of these samples is a promising possibility for determining the structure of the Co layers. As a general trend, however, further efforts to study structure and magnetic behavior should perhaps focus more on sample fabrication. If sample structure can be more easily controlled during sample manufacture, then IF 125 relationships between the structure and magnetic behavior may be more easily determined. 8.4.2 Fe/Si and Fe/{FeSi} Multilayers Further information is needed about both the structure and the magnetic state of the Fe in the {FeSi} spacer layers. By using isotopically-enriched Fe in the {FeSi} spacer layers, Mossbauer spectroscopy studies could be used to determine the magnetic state of Fe in the spacer and to infer its local structural environment since non-magnetic Fe and magnetic Fe have very different Mossbauer spectra. Since these spectra are sums of individual contributions, the measured data could be fitted using well-documented results for silicides and Fe—Si alloys to perhaps determine the actual composition of the spacer layer. lNoboru Sato, J. Appl. Phys. 63, 3476 (1988). 2Y. Wang, F.Z. Cui, W.Z. Li, and Y.D. Fan, J. Magn. Magn. Mater. 102, 121 (1991). "I11111111111111“1111'“