Top Banner
Nuclear magnetic ordering in simple metals at positive and negative nanokelvin temperatures A. S. Oja* and O. V. Lounasmaa Low Temperature Laboratory, Helsinki University of Technology, FIN-02150 Espoo, Finland This paper is a comprehensive review of almost twenty years of research on nuclear magnetic ordering, first in copper and later in silver and rhodium metals. The basic principles of nuclear magnetism and the measurement of positive and negative spin temperatures are discussed first. Cascade nuclear refrigeration techniques, susceptibility and nuclear-magnetic-resonance (NMR) measurements, and arrangements for neutron-diffraction experiments at nanokelvin and picokelvin temperatures are described next. Comprehensive magnetic-susceptibility and neutron-diffraction measurements on copper, which led to the discovery of at least three antiferromagnetic phases, one displaying the novel (0 2 3 2 3 ) spin structure and the other two showing the type-I order of the fcc system, are then described in detail. NMR data on silver, at T.0 and T,0, are presented next leading to the observation that silver orders antiferromagnetically at positive spin temperatures and ferromagnetically at negative spin temperatures. The authors discuss recent neutron-diffraction measurements that show that the antiferromagnetic structure at T.0 is in a single-k type-I state. NMR data on rhodium at T.0 and T,0 are also described. Results obtained on Tl, Sc, AuIn 2 , and metallic Pr compounds and on insulators like CaF 2 are then discussed briefly. The paper is concluded by an extensive theoretical section. Calculations of conduction-electron mediated exchange interactions are described, and the mean-field theory of nuclear magnetic ordering is presented. The role of thermal and quantum fluctuations is then discussed, particularly in the selection of the antiferromagnetic ground state. Finally, theoretically calculated magnetic phase diagrams and ordered spin structures of copper and silver are presented in detail and compared with experimental results. The overall agreement is good, affirming the value of nuclear magnets in Cu and Ag as realizations of simple physical models. [S0034-6861(97)00301-2] CONTENTS List of Symbols 3 I. Introduction 3 A. Historical comments 3 B. Theoretical survey 5 C. Susceptibility and nuclear-magnetic-resonance (NMR) experiments on copper 6 D. Neutron-diffraction measurements on copper 6 E. Experiments on silver and rhodium 8 F. Studies of other metals 9 II. Basic Principles of Positive and Negative Spin Temperatures in Nuclear Magnetism 10 A. Spin temperature 10 1. Zeeman temperature 10 2. Temperature of the interaction reservoir 10 B. Heat reservoirs, spin-lattice relaxation, and thermal mixing 10 C. Demagnetization to the ordered state 12 D. Negative spin temperatures 13 1. Production of negative temperatures 13 2. Thermodynamics at T,0 14 3. Nuclear ordering at T,0 15 E. Thermometry with nuclear spins 16 III. Experimental Techniques for Susceptibility Measurements 16 A. Principle of brute force nuclear cooling 16 B. Helsinki cryostat 17 C. Experimental procedure 19 D. Sample preparation 21 1. Quenching of magnetic impurities by internal oxidation 21 2. Thermal contact 22 E. SQUID measurements of the NMR signal and the low-frequency magnetic susceptibility 23 IV. Measurement of Nuclear-Spin Temperature 25 A. Calibration of polarization and entropy 25 B. Calibration of susceptibility 26 C. Local field 26 D. Consistency checks using high-T expansions 27 E. Secondary thermometers 28 F. Nonadiabaticities 28 G. Production of negative spin temperatures by a rapid field reversal 30 V. Susceptibility and NMR Data on Copper 31 A. Susceptibility and entropy at B50 31 B. Thermometry 33 C. Metastability and nonadiabatic phenomena 34 D. Field effects on a polycrystalline sample 34 E. NMR lineshapes of the ordered spin structures 35 F. Susceptibility data of a single-crystal specimen 36 G. Magnetization 38 H. Entropy diagram 38 VI. Experimental Techniques of Neutron-Diffraction Measurements 38 A. Setup for neutron-diffraction experiments 38 B. Risø cryostat 39 C. Beam heating 41 D. Use of polarized neutrons 42 1. Experimental setup 43 2. Flipping ratio versus polarization in copper 43 VII. Neutron-Diffraction Experiments on Copper 44 A. Neutron scattering from ordered copper nuclei 44 *Present address: VTT Automation, Measurements Technol- ogy, P.O. Box 1304, FIN-02044 VTT, Finland. Fax: 358-9-451- 2969. Electronic address: Aarne.Oja@vtt.fi 1 Reviews of Modern Physics, Vol. 69, No. 1, January 1997 0034-6861/97/69(1)/1(136)/$30.40 © 1997 The American Physical Society
136

Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

Aug 12, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

Nuclear magnetic ordering in simple metals at positiveand negative nanokelvin temperatures

A. S. Oja* and O. V. Lounasmaa

Low Temperature Laboratory, Helsinki University of Technology,FIN-02150 Espoo, Finland

This paper is a comprehensive review of almost twenty years of research on nuclear magneticordering, first in copper and later in silver and rhodium metals. The basic principles of nuclearmagnetism and the measurement of positive and negative spin temperatures are discussed first.Cascade nuclear refrigeration techniques, susceptibility and nuclear-magnetic-resonance (NMR)measurements, and arrangements for neutron-diffraction experiments at nanokelvin and picokelvintemperatures are described next. Comprehensive magnetic-susceptibility and neutron-diffractionmeasurements on copper, which led to the discovery of at least three antiferromagnetic phases, onedisplaying the novel (0 2

323) spin structure and the other two showing the type-I order of the fcc system,

are then described in detail. NMR data on silver, at T.0 and T,0, are presented next leading tothe observation that silver orders antiferromagnetically at positive spin temperatures andferromagnetically at negative spin temperatures. The authors discuss recent neutron-diffractionmeasurements that show that the antiferromagnetic structure at T.0 is in a single-k type-I state.NMR data on rhodium at T.0 and T,0 are also described. Results obtained on Tl, Sc, AuIn2, andmetallic Pr compounds and on insulators like CaF2 are then discussed briefly. The paper is concludedby an extensive theoretical section. Calculations of conduction-electron mediated exchangeinteractions are described, and the mean-field theory of nuclear magnetic ordering is presented. Therole of thermal and quantum fluctuations is then discussed, particularly in the selection of theantiferromagnetic ground state. Finally, theoretically calculated magnetic phase diagrams and orderedspin structures of copper and silver are presented in detail and compared with experimental results.The overall agreement is good, affirming the value of nuclear magnets in Cu and Ag as realizations ofsimple physical models. [S0034-6861(97)00301-2]

CONTENTS

List of Symbols 3I. Introduction 3

A. Historical comments 3B. Theoretical survey 5C. Susceptibility and nuclear-magnetic-resonance

(NMR) experiments on copper 6D. Neutron-diffraction measurements on copper 6E. Experiments on silver and rhodium 8F. Studies of other metals 9

II. Basic Principles of Positive and Negative SpinTemperatures in Nuclear Magnetism 10A. Spin temperature 10

1. Zeeman temperature 102. Temperature of the interaction reservoir 10

B. Heat reservoirs, spin-lattice relaxation, andthermal mixing 10

C. Demagnetization to the ordered state 12D. Negative spin temperatures 13

1. Production of negative temperatures 132. Thermodynamics at T,0 143. Nuclear ordering at T,0 15

E. Thermometry with nuclear spins 16III. Experimental Techniques for Susceptibility

Measurements 16A. Principle of brute force nuclear cooling 16B. Helsinki cryostat 17C. Experimental procedure 19

*Present address: VTT Automation, Measurements Technol-ogy, P.O. Box 1304, FIN-02044 VTT, Finland. Fax: 358-9-451-2969. Electronic address: [email protected]

Reviews of Modern Physics, Vol. 69, No. 1, January 1997 0034-6861/97/

D. Sample preparation 211. Quenching of magnetic impurities by internal

oxidation 212. Thermal contact 22

E. SQUID measurements of the NMR signal andthe low-frequency magnetic susceptibility 23

IV. Measurement of Nuclear-Spin Temperature 25A. Calibration of polarization and entropy 25B. Calibration of susceptibility 26C. Local field 26D. Consistency checks using high-T expansions 27E. Secondary thermometers 28F. Nonadiabaticities 28G. Production of negative spin temperatures by a

rapid field reversal 30V. Susceptibility and NMR Data on Copper 31

A. Susceptibility and entropy at B50 31B. Thermometry 33C. Metastability and nonadiabatic phenomena 34D. Field effects on a polycrystalline sample 34E. NMR lineshapes of the ordered spin structures 35F. Susceptibility data of a single-crystal specimen 36G. Magnetization 38H. Entropy diagram 38

VI. Experimental Techniques of Neutron-DiffractionMeasurements 38

A. Setup for neutron-diffraction experiments 38B. Risø cryostat 39C. Beam heating 41D. Use of polarized neutrons 42

1. Experimental setup 432. Flipping ratio versus polarization in copper 43

VII. Neutron-Diffraction Experimentson Copper 44

A. Neutron scattering from ordered copper nuclei 44

169(1)/1(136)/$30.40 © 1997 The American Physical Society

Page 2: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

2 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

B. Observation of nuclear magnetic ordering 45C. Magnetic-field dependence of the (1 0 0)

reflection for Bi[01 1] 45D. (0 2

323) reflection 47

E. Neutron-intensity contour diagram 49F. Kinetics of phase transitions when B i [01 1] 49

1. Initial temporal development of neutronintensity 49

2. Hysteresis at phase boundaries 493. Temporal changes in the width of the Bragg

peak 504. Decay of metastable states 505. Entropy losses 51

G. Comparison with simultaneous susceptibilitymeasurements 51

H. Other field directions 521. [100] directions 522. [110] directions 543. [111] directions 554. Selection of stable domains by external-field

alignment 55I. Intensity diagram for different field alignments 56J. Search for other antiferromagnetic Bragg peaks 58

VIII. Susceptibility and NMR Data on Silver 59A. Introduction to experiments at negative spin

temperatures 59B. Nuclear magnetic susceptibility of silver 60C. Nuclear ordering of silver at T.0 62

1. Search for the transition 622. Magnetic preparation of the sample by the

demagnetization scheme 633. Antiferromagnetic phase boundary 644. Number of antiferromagnetic phases 655. Comparison with theory 65

D. Nuclear ordering in silver at T,0 661. Observation of the ferromagnetic transition 662. Phase boundary of the domain state 673. Comparison with theory 68

IX. Neutron-Diffraction Experiments on Silver 68A. Experimental arrangements 68B. Results 69C. Comparisons with theory 70

X. Studies on Rhodium 71A. Nuclear magnetic susceptibility 71B. Exchange constants and ordered spin structures 73C. Spin-lattice relaxation in rhodium 74

XI. NMR Studies in the Paramagnetic Phase 76A. Cross relaxation between 107Ag and 109Ag 76

1. Measurement of the cross-relaxation time tx 772. Theoretical description 783. Extraction of the exchange constant 79

B. Polarization-induced suppression andenhancement of isotopic NMR lines 801. Copper 802. Silver 803. Theoretical calculations 81

C. Second-harmonic Larmor line 81XII. Experiments on Thallium, Scandium, and AuIn2 82

A. Thallium 82B. Scandium 82C. Nuclear ordering in AuIn2 84

XIII. Hyperfine-Enhanced Nuclear Magnetism inPraseodymium Compounds 85A. Nuclear refrigeration 86B. Nuclear ordering 87

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

XIV. Spontaneous Nuclear Order in Insulators at T.0and T,0 88A. Dynamic nuclear polarization 88B. Adiabatic demagnetization in the rotating frame 89C. Truncated dipolar Hamiltonian 90D. Experimental results on CaF2 , LiH, and

Ca(OH)2 90XV. Theory 91

A. Exchange interactions 911. Ruderman-Kittel interaction 922. Anisotropic exchange interactions 933. Other interactions 94

B. Mean-field theory of magnetic ordering 951. Basic equations 952. Eigenvalue and other k-space equations 953. Ordering vector and the ordering temperature 954. Equal-moment and permanent spin structures 965. Thermodynamics 966. Transition from the polarized paramagnetic

state to the antiferromagnetic phase 967. Ordering in the fcc lattice 97

C. Comparison of measured and calculatedmagnetic properties of Cu, Ag, and Rh 981. Mean-field Tc and the ordering vector 982. Bc(T50) and x of the ordered state 1003. Beyond the mean-field theory 101

a. Spherical model 101b. High-T expansions and other quantum-

spin theories 102c. Monte Carlo simulations 102

4. S and x of the paramagnetic state 103D. Fluctuation-stabilized type-I spin configurations 103

1. Continuous degeneracy of the mean-fieldsolution 103

2. Static susceptibility matrices 1043. Overview of fluctuation mechanisms 1044. Thermal fluctuations 1055. Quantum fluctuations 1056. Isotropic spin-spin interactions 1067. Anisotropic spin-spin interactions with an

easy plane 107a. Bi[001] 108b. Bi[110] 109c. Bi[111] 111

8. Comparison with experiments 111E. NMR response of type-I structures 112

1. Equations of motion 1122. Resonances 1123. Simulation of spin dynamics 114

F. (0 23

23) order 115

1. Principal features of the (0 23

23) spin

configurations 115a. Easy-axis anisotropy 115b. Up-up-down structure 115c. Superposition with type-I order 115d. Construction of multiple-k structures 116

2. Theoretical spin structure versus experimentson copper 116a. Bi[011] 117b. Bi[100] 118c. Bi[111] 119d. Other field directions 119

3. Stability of (0 23

23) modulation versus type-I

order 1194. Superposition structure or a mixed state of

(0 23

23) and (1 0 0) domains 120

Page 3: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

3A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

5. Related electronic magnets 121G. High-field phase of copper when Bi[111] 121

1. A (h k l) structure 121a. Model Hamiltonian 121b. Interplay of the (h k l), (0 2

323), and

(1 0 0) modulations 122c. Criticism 124

2. Other possibilities 124H. Ferromagnetic ordering at T,0 125

1. Domain configurations 1252. Demagnetization into the domain state 1263. Comparison with experimental data on silver 127

XVI. Summary and Future Prospects 127A. Copper 127B. Other simple metals 127C. Temperature records 128

Acknowledgments 128References 129

LIST OF SYMBOLS

Most of the specialized notation is defined where it isneeded in the text. This list collects for ready referencethe notation used throughout the paper.

B loc The average strength of the local fluctuating field[Eq. (29)].

Bi Molecular field at site i due to interactions withneighboring spins [Eq. (75)].

di Amplitude of the antiferromagnetic modulationcorresponding to the ordering vector ki [see, forexample, Eqs. (111) and (160)].

D Demagnetizing factor. It is assumed that thedemagnetizing tensor is diagonal and that D isthe component of interest, usually the oneparallel to the external magnetic field.

L Lorentz constant L513 [see, for example, Eq.

(34)].p Nuclear-spin polarization p5^Ii&/I [Eq. (76)].R Strength of the exchange interaction with respect

to the dipolar force [Eqs. (6) and (66)].T Temperature of nuclear spins.Te Temperature of conduction electrons.TZ Temperature of the Zeeman reservoir (see Sec.

II.B.).Tss Temperature of the spin-spin interaction

reservoir (see Sec. II.B).ln(k) Eigenvalue, or eigenenergy, of the Fourier

transform of the interaction matrix A(k) [seeEq. (83)].

lmax The largest eigenvalue lmax5maxk,n$ln(k)%,sometimes denoted as l. See Eqs. (87) and (88).

u Weiss temperature u5C(R1L2D) in x5C/(T2u) [see Eq. (34)].

I. INTRODUCTION

This paper reviews two decades of research onnuclear magnetic ordering, first in copper and later insilver and rhodium metals. The focus, in the experimen-tal section, is on research carried out by the authors andtheir colleagues in Helsinki, as well as on joint projects

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

in the Riso” National Laboratory (Denmark) and at theHahn-Meitner-Institut (Berlin), in the course of whichat least three antiferromagnetic phases were discoveredand several low-temperature world records broken atboth positive and negative sides of the absolute zero.This work is set in the context of the broader effort tounderstand spin structure and to measure magneticphases in a variety of metals, metallic compounds, andinsulators.

Nuclear spin assemblies in simple metals are goodmodel systems of magnetism, a fact that has motivated alarge body of theoretical work. We review this work inthe last third of the paper. Our emphasis is on copper,silver, and rhodium, particularly on copper, in part be-cause of our own experience with these systems and inpart as a reflection of the huge body of experimental andtheoretical work that has been produced on these met-als.

The review is organized as follows. In Sec. I.A. webriefly survey the historical background and in I.B thetheory of nuclear ordering studies. Section I.C. discussesthe early susceptibility measurements and NMR experi-ments using continuous-wave techniques. Neutron-diffraction measurements on copper followed (Sec. I.D),and then experiments on silver and rhodium (Sec. I.E)and other metals (Sec. I.F). Section II describes the prin-ciples of nuclear magnetism that are important in thestudy of metals. This completes the background portionof the paper.

Three experimental techniques are treated next: sus-ceptibility measurements (Sec. III), measurement ofnuclear-spin temperatures (Sec. IV), and neutron-diffraction techniques (Sec. VI). The data obtained bythese methods on copper are discussed in Secs. V andVII. Results on silver (Secs. VIII and IX) and rhodium(Sec. X) are then presented. Other NMR studies of cop-per and silver, not directly related to nuclear ordering,are described in Sec. XI. Recent nuclear ordering ex-periments on other elements and materials such as thal-lium, scandium, gold-indium, praseodymium com-pounds, and insulators are reviewed in Secs. XII–XIV.

The longest single section, amounting to about one-third of the review, is Sec. XV, which treats in detail thetheoretical work on spontaneous nuclear ordering insimple metals. Its length is also an indication of the ap-peal that the subject holds for researchers. See the be-ginning of Sec. XV for a more detailed description of itscontents.

In Sec. XVI we summarize what has been learned todate about magnetic ordering in the three best-studiedsimple metals, copper, silver, and rhodium, giving mag-netic phase diagrams for each. We also mention somepossible directions that future studies of nuclear order-ing could take.

A. Historical comments

Electronic magnetism exhibits a wide spectrum of dif-ferent ordering phenomena over a temperature rangethat extends from room temperature and above in iron

Page 4: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

4 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

to a few millikelvins in cerium magnesium nitrate. Be-cause nuclear magnetic moments are three orders ofmagnitude smaller than their electronic counterpartsand because the interactions between spins are propor-tional to the magnetic moment squared, spontaneous or-dering phenomena can be expected to occur in thenuclear-spin system only at microkelvin temperaturesand below. Solid 3He is an exception owing to thestrong quantum-mechanical exchange force, enhancedby the large zero-point motion (Roger et al., 1983; Crossand Fisher, 1985), and so are Van Vleck paramagnets,such as PrNi 5, in which considerable hyperfine enhance-ment of the magnetic field at the nucleus occurs (Andresand Lounasmaa, 1982).

Copper is probably the best material for studies ofnuclear ordering, since cooling by adiabatic nuclear de-magnetization of this metal is relatively easy owing to itsfavorable thermal and magnetic properties. Progress incooling techniques has recently made it possible to carryout similar experiments on silver and rhodium. The op-portunity to make detailed comparisons with theoreticalpredictions is the most important driving force for ex-periments on spontaneous nuclear ordering in simplemetals. Such metals offer several theoretical advantages:the spins are well localized, the magnetic degrees offreedom do not couple to lattice distortions, and thespin-spin interactions are well defined. Simple metalscan be exchange-dominated magnets, such as silver andrhodium, or magnets with equally strong dipolar and ex-change forces, such as copper.

The pioneering experiments of Nicholas Kurti and co-workers established, in 1956, the feasibility of thenuclear-demagnetization method (Kurti et al., 1956;Hobden and Kurti, 1959). In spite of the limitations im-posed by cryogenic techniques available at that time, theOxford group succeeded in reaching 1 mK. The firststudies of nuclear cooperative phenomena, on insulatorslike CaF 2 and LiH, were made by Abragam and Gold-man and co-workers at Saclay 13 years later (Chapellieret al., 1969; Abragam and Goldman, 1982). The Helsinkiwork on copper (Ehnholm et al., 1980; Huiku et al.,1986, and references therein) was begun as early as 1974but, due to many experimental difficulties, only since thespring of 1982 has spontaneous nuclear order been ob-served below the Neel temperature TN558 nK. As aculmination point of this research, in 1984 the magneticfield vs entropy diagram (see Fig. 1) of the nuclear-spinsystem in copper was constructed, with the external fieldin the crystallographic [001] direction of the single-crystal sample; three antiferromagnetic phases werefound.

In 1987, a Danish-Finnish-German group, working atthe Risø National Laboratory near Copenhagen, suc-ceeded in observing, by means of neutron scattering, the(1 0 0) Bragg reflection from a nuclear-ordered single-crystal 65Cu specimen (Jyrkkio, Huiku, Lounasmaa,et al., 1988). This study resolved the spin order in two ofthe antiferromagnetic phases. In 1989, the nature of the

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

third phase was determined by observing a strong neu-tron peak at the (0 2

323) Bragg position (Annila et al.,

1990).Experiments on silver were begun in Helsinki in 1987.

Because of the relatively small magnetic moment of theAg nuclei, an ordering temperature much lower thanthat observed in copper was expected. The NMR mea-surements indicated a transition to an antiferromagneticstate at TN5560 pK. The long spin-spin relaxation timeof silver made it possible to produce negative nuclear-spin temperatures and look for nuclear ordering atT,0. The measurements showed that the same interac-tions that produced antiferromagnetism at positive tem-peratures ordered the nuclear-spin system of silver fer-romagnetically above TC5 –1.9 nK. These experimentshave been described in several papers (see, for example,Oja, Annila, and Takano, 1990; Hakonen and Yin, 1991;Hakonen, Nummila, Vuorinen, and Lounasmaa, 1992;Steiner, Metz, Siemensmeyer, et al., 1996).

Most recently the Helsinki group has investigatednuclear ordering in rhodium (Hakonen, Vuorinen, andMartikainen, 1993). The spin-spin interactions in thismetal are slightly weaker than in silver. So far rhodiumnuclei have been cooled to 280 pK; this is the currentworld low-temperature record. In the negative tempera-ture regime, T5 –750 pK has been reached, which is, ina sense, the highest temperature obtained so far. That is,the energy of a spin system is larger at a negative tem-perature than at, say, an infinite positive temperature(see Sec. II.D). In spite of these records, nuclear order-ing in rhodium has not yet been observed.

FIG. 1. External magnetic field vs entropy diagram for nuclearspins in copper. The three antiferromagnetic phases are de-noted by AF1, AF2, and AF3; P stands for the paramagneticphase and the shaded areas indicate regions where first-orderphase transitions take place. The upper right corner of thefigure shows the schematic spin arrangements that were origi-nally proposed. The suggested structure for AF2, which con-sists of two sublattices, is, however, inconsistent with laterneutron-diffraction experiments (see Fig. 3). In addition, AF1and AF2 can, in fact, consist of four sublattices rather thantwo. From Huiku et al. (1986).

Page 5: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

5A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

B. Theoretical survey

At very low temperatures it is meaningful to speakabout two distinct temperatures in the same specimen,at the same time: the nuclear-spin temperature, denotedin this paper by T , and the common conduction-electronand lattice temperature Te . The nuclei reach local ther-mal equilibrium among themselves in a time character-ized by t2, the spin-spin relaxation time (150 ms in cop-per and 10 ms in silver and in rhodium), whereas theapproach to equilibrium between nuclear spins and con-duction electrons is governed by the spin-lattice relax-ation time t1. At low temperatures, t2!t1; this makes aseparate nuclear-spin temperature meaningful (Gold-man, 1970). In experiments on copper, silver, andrhodium, Te and T may differ by several orders of mag-nitude. Nevertheless, thermal isolation between the con-duction electrons and the nuclei is sufficient for the spinsystem to be treated by equilibrium thermodynamics.

Nuclear ordering can be studied by using several cool-ing techniques, described in many review papers andmonographs (Huiskamp and Lounasmaa, 1973; Lounas-maa, 1974; Betts, 1976; Abragam and Goldman, 1982;Andres and Lounasmaa, 1982; Pickett, 1988; Pobell,1992a, 1992b). Among them, the ‘‘brute force’’ method,applicable to metallic samples, is the most straightfor-ward. In this technique, the nuclear spins are first polar-ized using a high initial magnetic field Bi (6–9 T) at alow starting temperature Ti (10–20 mK) and then adia-batically demagnetized to a small final field Bf . Underideal conditions, the temperature then decreases accord-ing to the relation Tf5TiBf /Bi , provided that Bf isclearly larger than the local internal field of the material.Further cooling can be achieved by cascading twonuclear-demagnetization stages. If demagnetization isstarted from a sufficienly low spin entropy, and if Bf issmall enough, it can be expected that spin-spin interac-tions will produce spontaneous magnetic order in theassembly of nuclei. This ordering is analogous to phe-nomena observed in electronic systems at much highertemperatures. All investigations of nuclear magnetism incopper, silver, and rhodium have been carried out usingthe brute force cascade-demagnetization technique. Itshould be noted that nuclear-spin order is not influencedby the temperature of the conduction electrons.

This review is devoted to experiments on and theoryof nuclear ordering in Cu, Ag, and Rh. Data on thal-lium, scandium, AuIn 2, and praseodymium compoundswill be discussed only briefly. In Cu, Ag, and Rh the spinHamiltonian is

H5HD1HRK1HZ . (1)

The first term is the dipole-dipole interaction

HD5~m0\2/4p!(i,j

g ig jr ij23@Ii•Ij23rij

22~Ii•rij!~Ij•rij!# ,

(2)

where m0 is the permeability of free space, g i is the gy-romagnetic ratio of the nucleus at the lattice site i , rij isthe lattice vector from i to j , Ii is the nuclear spin, and

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

the summation is to be taken over all pairs i ,j . The sec-ond term in Eq. (1) is the Ruderman-Kittel (RK) cou-pling (Ruderman and Kittel, 1954)

HRK52(i,j

J ijIi•Ij , (3)

where Jij is the coupling constant between spins at sitesi and j . The RK interaction is usually the largest contri-bution to the indirect exchange force. In copper, aniso-tropic exchange is also significant although clearlysmaller than the RK force (see Sec. XV.A.2). The lastcontribution in the Hamiltonian of Eq. (1) is the Zee-man term

HZ52\gB•(i

Ii , (4)

where B is the applied magnetic field.While the dipolar and Zeeman forces are accurately

known, exchange interactions must be inferred eitherfrom experiments or from theoretical calculations. Inthe free-electron approximation, Jij assumes the form

Jij5h~m0/4p!\2g2rij23f~2kFrij! , (5)

where f(x)5cos(x)2sin(x)/x and the parameter h de-scribes the strength of the RK interaction. Theoreticalvalues for Jij in copper and silver, more reliable thanthose obtained using the free-electron approximation,have been deduced recently from band-structure calcu-lations (Lindgard et al., 1986; Frisken and Miller, 1986,1988b; Miller and Frisken, 1988; Oja et al., 1989; Har-mon et al., 1992).

Experimentally, the relative magnitudes of the RK in-teraction and the dipolar forces are often described bythe parameter

R5(j

J ij /~m0\2g2r! , (6)

where r is the number density of spins. The R parameterhas the value −0.42 for copper, −2.5 for silver, and –1.4for rhodium. In an fcc lattice, with one conduction elec-tron per atom, the corresponding h parameters in Eq.(5) are −0.71, −4.3, and −2.4, respectively (R50.587h).The negative signs of R and h imply that the interac-tions have an antiferromagnetic character. Thus we seethat copper is in the region where the dipolar and RKinteractions are of comparable strength, while silver andrhodium are exchange dominated.

The gyromagnetic ratios of the two stable isotopes ofthese elements, 63Cu and 65Cu in copper and 107Ag and109Ag in silver, differ by 7% and 13%, respectively.Rhodium has only one stable isotope, 103Rh. The vari-ous g’s are given in Table I. There are no nuclear qua-drupolar interactions in silver or rhodium because allstable isotopes have a spin I5 1

2. Neither is there a qua-drupolar force in an undistorted lattice of copper(I5 3

2) owing to fcc symmetry. The Hamiltonian of Eq.(1) should thus describe the nuclear-spin systems inthese metals rather well in a carefully prepared speci-men.

Page 6: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

6 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

C. Susceptibility and nuclear-magnetic-resonance (NMR)experiments on copper

The main nuclear ordering results on copper were ob-tained either by means of ac susceptibility measure-ments at low frequency or by NMR experiments usingcontinuous-wave techniques. The NMR studies were un-usual in that the resonant frequencies were low becausethe measurements were made either in zero field or in asmall external field. The interesting range of frequencieswas typically from 0.5 to 20 kHz in copper and evenlower in silver and rhodium, from 10 to 200 Hz. In prin-ciple, the NMR experiment was just an ac susceptibilitymeasurement in which the response of the spin systemwas monitored in two components, separated by a 90-degree phase angle, and the frequency was swept acrossthe maximum of the absorption signal to observe theresonance. The static susceptibility x8(0) was calculatedfrom the measured NMR absorption peak by using theKramers-Kronig relation. The ac susceptibility experi-ments, at a sufficiently low frequency, give x8(0) quiteaccurately, too. Therefore, in the terminology of this re-view, static susceptibility data on copper were, in fact,the results of low-frequency ac susceptibility measure-ments.

The first NMR results on copper were obtained in1978 (Ehnholm, Ekstrom, Jacquinot et al., 1979; Ek-strom et al., 1979): The nuclear spins reached about 50nK in zero field, but no clear evidence for a magneticphase transition was seen (Kurti, 1982). However, sev-eral important results were deduced, among them themagnitude and sign of the RK exchange interaction. Thevalue R=−0.42 was obtained in two ways: from the rela-tive magnitudes of the separate 63Cu and 65Cu NMRabsorption lines and from a 2-kHz shift downwards ofthe second harmonic in the NMR peak. According to acalculation based on the mean-field theory (Kjaldmanand Kurkijarvi, 1979), the RK interaction with R=−0.42should have ordered copper nuclei to an antiferromag-netic state as early as 230 nK, far above the lowest tem-peratures reached in the measurements. This discrep-ancy between experiment and theory remained aproblem for several years, although calculations showedthat fluctuations can substantially decrease the orderingtemperature from its mean-field value (Kumar et al.,1980; Kjaldman et al., 1981; Niskanen and Kurkijarvi,1981, 1983; Niskanen et al., 1982).

The main problem in these early experiments was theshort spin-lattice relaxation time, about 10 minutes,which caused the entropy of the nuclear-spin system to

TABLE I. Properties of stable Cu, Ag, and Rh isotopes, im-portant for nuclear ordering experiments.

Isotopic I g/2p

abundance (%) (kHz/mT)

63Cu, 65Cu 69.1, 30.9 32, 3

2 11.29, 12.09107Ag, 109Ag 51.8, 48.2 1

2, 12 1.723, 1.981

103Rh 100 12 1.340

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

increase rapidly after demagnetization. In the next seriesof measurements (Ehnholm et al., 1980), started in 1979,Te was decreased further, which resulted in a somewhatlonger relaxation time. It was then observed, after de-magnetization to zero field, that the susceptibility wassaturated during the first few minutes in a longitudinalmeasuring geometry but not in a transverse geometry.Subsequent studies suggested that the plateau in the lon-gitudinal susceptibility had been caused by an experi-mental artifact (Huiku and Soini, 1983). However, anunexplained difference between the longitudinal andtransverse zero-field susceptibilities remained.

Finally, in 1982, the Helsinki group unambiguouslysaw ordering in copper at Tc558 nK (Huiku andLoponen, 1982; Huiku, Jyrkkio, and Loponen, 1983); thenuclear-spin system was cooled to the at-that-timerecord low temperature of 25 nK. A polycrystallinespecimen of thin foils was used. Two critical modifica-tions finally led to experimental success: the preparationof the sample by selective oxidation and the construc-tion of a new nuclear cooling stage made of a bulk pieceof copper. These improvements resulted in an order-of-magnitude increase of the spin-lattice relaxation time.Simultaneously, indications of two different ordered re-gions were found by recording the static susceptibilityx8(0) during warmup after demagnetization to differentfinal fields below the critical value Bc50.25 mT.

Experiments on a single-crystal specimen were begunin 1984 (Huiku, Jyrkkio, et al., 1984; Huiku et al., 1986).The static susceptibility was measured in the three Car-tesian directions. Below Bc , three different antiferro-magnetically ordered regions were distinguished at lowentropies. Using a simple two-sublattice model for thetransverse and longitudinal susceptibilities, provisionalspin arrangements were obtained for the ordered states.The B-S phase diagram, illustrated in Fig. 1, was con-structed.

The experimental result was analyzed in several theo-retical calculations (Kjaldman and Kurkijarvi, 1979; Ojaand Kumar, 1984; Kumar et al., 1985, 1986; Viertio andOja, 1987; Frisken and Miller, 1988a; Lindgard, 1988a).However, it was not possible to test the proposed spinstructures against experiments, since measurements ofstatic susceptibility do not yield sufficiently detailed in-formation. In studies of electronic magnetism, neutrondiffraction has been the tool for determining orderedspin structures. The basic principles of nonmagnetic andmagnetic Bragg reflections are illustrated in Fig. 2. Inthe case of nuclear magnets, the use of neutron diffrac-tion is based on the spin-dependent part of the neutron-nucleus scattering amplitude, which results from thestrong interaction between a neutron and a nucleus. Inparticular, long-range antiferromagnetic order gives riseto additional Bragg reflections, yielding directly thetranslational symmetry of the spin system.

D. Neutron-diffraction measurements on copper

Previously, the ordered structures of purely nuclearsystems had been determined by neutrons in only a few

Page 7: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

7A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

cases and never in simple metals. The first measure-ments were made on LiH in 1978 by Abragam andGoldman and their co-workers at Saclay (Roinel et al.,1978; Abragam and Goldman, 1982).1 Neutron studiesof this insulator are feasible because of the very largespin-dependent scattering amplitude of protons. Theseexperiments revealed antiferromagnetic and ferromag-netic nuclear structures at T.0 and T,0, respectively.The ordering temperatures could not be determinedfrom the experimental data. Benoit et al. (1985) wereable to perform neutron-diffraction experiments on or-dered solid 3He and to verify the previously proposed(Osheroff et al., 1980) up-up-down-down order in thelow-field phase. Even though the transition temperatureis as high as 1 mK, other experimental difficulties, suchas the need to grow the 3He crystal in situ and the enor-mous neutron absorption by the 3He nuclei, kept thecounting statistics rather low. Nuclear-spin ordering hasalso been observed around 1 mK by neutron diffractionin several hyperfine-enhanced systems (Benoit et al.,1981; Nicklow et al., 1985).

Before the Risø work on copper was begun in 1986(Jyrkkio, Huiku, Clausen, et al., 1988), neutron-diffraction studies had not been performed on a purelynuclear magnet with a full dipolar and exchange interac-tion between the spins. As metallic copper provides sucha system, the study paved the way for an understandingof ordering in this metal as well as in nuclear magnetsmore generally. The large amount of information ob-tained earlier by the susceptibility measurements in Hel-sinki provided a good basis for subsequent investigationswith neutrons. The Risø study was performed on an iso-topically enriched 65Cu single crystal. As the orderingtook place at 58 nK, these measurements opened a to-tally new temperature range for neutron-diffraction ex-periments.

Reaching low nanokelvin temperatures in quiet sur-roundings is a difficult undertaking in any case, but thetask is even more complicated in a reactor environment.The presence of numerous pieces of equipment in theexperimental hall tends to increase the heat leak into thecryogenic system. A new and significant problem iswarming of the specimen by the neutron beam. The low-temperature requirements must be taken into account inthe design of the diffractometer as well. A special fea-ture, not encountered in standard neutron-scattering ex-periments, is that only limited mechanical movements ofthe spectrometer are allowed during actual measure-ments, again because of heating problems.

1In the studies of insulators by the Saclay group, nuclear or-der was produced in the rotating frame of reference, under ahigh magnetic field, by the so-called truncated part of the di-polar interaction, i.e., the part that commutes with the Zeemanterm (see Sec. XIV.C). For dynamic polarization, a small con-centration of paramagnetic impurities was introduced into thesample. The ordering temperatures in these insulators were inthe range 0.3–0.6 mK according to theoretical estimates.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

The Risø group enjoyed its first significant success inthe fall of 1987, when the antiferromagnetic (1 0 0)Bragg reflection from nuclear-ordered copper was de-tected (Jyrkkio, Huiku, Lounasmaa et al., 1988). The sig-nal corresponded to type-I ordering in the fcc lattice; theobservation was consistent with theoretical predictions(Kjaldman and Kurkijarvi, 1979). In the simplest case, atype-I antiferromagnetic spin configuration consists ofalternating ferromagnetic sheets. Such a structure is il-lustrated in Fig. 2(b).

FIG. 2. Illustration of basic principles of Bragg reflection. (a)k and k8 are the wave vectors of the incident and reflectedneutron beams, respectively. Q5k2k8 is the scattering vector.The scattering angle 2u is determined by the distance betweenthe lattice planes, as described by the equation at right. (b) Inthe presence of antiferromagnetism, there will be additionalBragg reflections resulting from the magnetic structure. For asimple two-sublattice structure, the fundamental antiferromag-netic reflection corresponds to an interplanar distance d52a ,where a is the lattice constant. Such a two-sublattice state pro-duces the antiferromagnetic (1 0 0) Bragg reflection in an fcc

lattice. (c) Illustration of the antiferromagnetic (0 23

23 ) Bragg

reflection in an fcc lattice. The spin configuration is the up-up-down pattern depicted in three dimensions in Fig. 3(c). Here,all spins have been projected onto the scattering plane, i.e., theplane spanned by k and k8. The period of the magnetic struc-ture d53a/A2.

Page 8: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

8 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Comprehensive neutron-diffraction measurementswere then made of the magnetic-field dependence of thescattered-neutron intensity (Jyrkkio et al., 1989). Thedata, shown in Fig. 3, indicated the existence of at leasttwo different antiferromagnetic phases, both producingthe (1 0 0) Bragg reflection. These spin structures werefound in the low- and high-field regions. In intermediatefields, the (1 0 0) reflection nearly vanished, and the na-ture of the spin structure there remained unclear.

In the fall of 1989 the puzzle was solved: anotherBragg reflection at (0 2

323) was discovered in the

intermediate-field region (Annila et al., 1990). This kindof ordering had not been observed before in any fccantiferromagnet. Once again, calculations were impor-tant in finding the antiferromagnetic reflection (Lind-gard et al., 1986; Lindgard, 1988a). Theoretical workidentified the (0 2

323) structure with an up-up-down spin

configuration in fields where the Bragg reflection wasstrongest (Viertio and Oja, 1990a). The spin structureand the resulting Bragg reflection are illustrated in Fig.2(c).

In the low intermediate-field region, both the (0 23

23)

and the (1 0 0) reflections were observed. Although thiscan be attributed to the presence of two different typesof domains, there is theoretical evidence for a more ex-otic type of ordering in which a single magnetic domaindisplays, at the same time, both reflections. The calcu-lated spin structures in the various regions of the B-Splane are illustrated in Fig. 3. The phase diagram ob-tained from the neutron-diffraction measurements is ingood overall agreement with the diagram constructedpreviously from magnetic-susceptibility measurements(see Fig. 1).

The most recent series of neutron-diffraction experi-ments at Risø (Annila et al., 1992) were performed byvarying the direction of the external magnetic field withrespect to the crystalline axis. It was found that the in-tensities of the (0 2

323) and the (1 0 0) Bragg reflections

depended sensitively on the alignment of the field.These measurements provided stringent tests of thetheoretically proposed spin structures. The data showed,indeed, the characteristic features of the predicted spinstructures (Viertio and Oja, 1987, 1990a, 1990b). Unex-pectedly, however, in high fields aligned close to the[111] crystalline axis, no (1 0 0) Bragg reflection was ob-served. It is likely that there is yet another type of anti-ferromagnetic order present in this field region, to befound in future studies.

E. Experiments on silver and rhodium

Measurements of nuclear magnetism in silver werebegun in Helsinki in 1986. The main difficulty in theseexperiments was achieving high enough initial nuclearpolarization, i.e., low enough spin entropy, before de-magnetizing the sample to zero field. The equilibriumpolarization at B57 T and at T5200 mK is 91.4% (thecorresponding entropy is 0.26R ln2, where R is the gasconstant), while in copper it would be 99.9%. Further-more, polarization of silver nuclei is a relatively slow

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

process: the spin-lattice relaxation time in a high mag-netic field is 14 h at Te5200 mK (t1 would be 2 h forCu).

In the first sets of experiments (Oja, Annila, and Ta-kano, 1990, 1991), the nuclear-spin entropy of the speci-men was reduced to 0.50R ln2, corresponding to polar-izations up to 78%. Evidence for antiferromagneticordering was obtained at the lowest entropies by observ-ing saturation of the static magnetic susceptibility. Con-clusive data on antiferromagnetic ordering were ob-tained, however, only after improved precooling of thesample. In this second set of experiments spin polariza-tions up to 94% were reached (Hakonen, Yin, and Lou-nasmaa, 1990; Hakonen and Yin, 1991; Hakonen, Yin,and Nummila, 1991; Hakonen, Nummila, and Vuorinen,1992; Hakonen, Nummila, Vuorinen et al., 1992; Ha-konen and Vuorinen, 1992). The ordering temperaturein zero field was determined to be TN5560 pK; this isthe lowest ordering temperature ever measured.

The boundary for antiferromagnetic order was tracedin the B-S plane. The phase diagram is illustrated in Fig.4. Interestingly, the critical entropy for ordering in-creased with B in small fields. Only one ordered statecould be singled out, although some features of theNMR spectra indicated the presence of two phases.Definite conclusions were difficult to make because thesample was polycrystalline.

An important new feature in the investigations on sil-ver was that it was possible to study nuclear magnetismat negative spin temperatures as well (Oja, Annila, andTakano, 1991). Unlike the spins in insulators studiedby the Saclay group (Abragam and Goldman, 1982),nuclear spins in silver interact through the full dipolarforce, which, together with exchange interactions, pro-duces ordering in the laboratory frame in low externalfields. There are several ways to generate negative spintemperatures. In silver this regime was achieved byquickly reversing the direction of the external magneticfield, in less than t2510 ms, so that the spins did nothave time to redistribute themselves among the energylevels. At negative temperatures, the population of spinsamong the Zeeman levels is given by the usual Boltz-mann factor, exp(2mB/kT) as at positive temperatures,but with T,0. Only the highest energy level is popu-lated when T→20. Therefore, the energy of a nuclear-spin system is maximized when absolute zero is ap-proached from the negative side. It can be shown that, atT,0, the thermodynamic equilibrium state correspondsto the maximum of the Gibbs free energy (see Sec.II.D.3).

Measurement of NMR absorption at T,0 is actuallya measurement of NMR emission, since x9(f),0. Staticsusceptibility can still be obtained using the Kramers-Kronig relation, that is, by integrating over x9(f)/f .Therefore, x8(0) is negative as well. Nuclear-spin tem-peratures can be determined by directly applying thesecond law of thermodynamics, using the same proce-dure as at T.0 (see Sec. IV). The x8(0) data showedthat silver orders ferromagnetically at T521.9 nK (Ha-konen, Nummila, Vuorinen, and Lounasmaa, 1992). Al-

Page 9: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

9A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

though it may, at first, seem surprising that a system withantiferromagnetic interactions can order ferromagneti-cally, this is exactly according to theoretical expectationsfor a spin system at T,0: Maximum energy for antifer-romagnetic interactions presumably corresponds to fer-romagnetic order. Such a situation, antiferromagnetismat T.0 and ferromagnetism at T,0, has also beenfound for LiH in a coordinate frame rotating at the Lar-mor frequency about the external magnetic field (Roinelet al., 1980). According to theoretical calculations (Vier-tio and Oja, 1992), the ferromagnetic structure in silveris a novel kind of domain configuration. The measuredvalue of the susceptibility in the ferromagnetic state isconsistent with the proposed structure. The phase dia-gram of silver at T,0 is illustrated in Fig. 4.

NMR measurements alone cannot verify details of thespin configuration in silver. Experiments employing scat-tering of polarized neutrons are again necessary for thispurpose. Tuoriniemi, Nummila et al. (1995) have re-cently observed long-range nuclear antiferromagneticorder by neutron diffraction in a single crystal of silver.The observed antiferromagnetic (0 0 1) Bragg peak de-cisively proved spontaneous long-range ordering. In amagnetic field along the [001] direction, a single-k statewith k5(p/a)(0,0,1) developed. A structure with thisordering vector remained stable in zero field. Domainswith the other equivalent k-vectors did not appearwithin the available experimental time, as was deducedfrom the lack of (0 0 1) intensity in measurements withBi[010]. The results support the theoretically predictedstructure for low magnetic fields along the [001] axis, butno evidence for a proposed triple-k state in higher fieldswas seen (Viertio, 1992; Heinila and Oja, 1993a). In 50mT ('Bc/2), the (0 0 1) reflection was observed when Bwas aligned within a cone of 110° opening around the[001] axis, in partial agreement with Monte Carlo simu-lations. The phase diagram based on neutron-diffractiondata closely resemble the earlier NMR results on a poly-crystalline sample of natural silver.

The spin dynamics of silver nuclei have been investi-gated extensively as well. Studies of cross relaxation(Oja, Annila, and Takano, 1990) between the Zeemantemperatures of the two isotopes 107Ag and 109Ag pro-vide a clear demonstration of the theoretical prediction(Goldman, 1970) that mutual flips of unlike spins, whichbecome effective once the two isotopic NMR lines over-lap, do not lead to an equalizing of temperature. Equi-librium between the two Zeeman reservoirs is achievedonly through the energy reservoir of spin-spin interac-tions via single spin flips, which become frequent in lowfields. The observed cross-relaxation rate also gives ameasure for the strength of the RK interaction in silver.Values for R were obtained from the relative intensitiesof the 107Ag and 109Ag NMR-absorption lines at a highspin polarization in a low field, both at T.0 and atT,0 (Hakonen, Nummila, and Vuorinen, 1992). Thedata were consistent with the early NMR work at hightemperatures (Poitrenaud and Winter, 1964), with sus-ceptibility measurements (Hakonen and Yin, 1991), andwith band-structure calculations (Harmon et al., 1992).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

During the past four years the Helsinki group has in-vestigated nuclear magnetism in rhodium metal, which isan I5 1

2 fcc system like silver (Hakonen, Vuorinen, andMartikainen, 1993, 1994). The R parameter [see Eq. (6)]was found to be –1.4, implying that interactions are ex-change dominated but not so strongly as in silver. Al-though the nuclear magnetic moment of rhodium is 30%smaller than in silver, nuclear-spin polarizations up to83% were reached in experiments at T.0 and –60% atT,0. These values correspond to spin entropies that insilver were low enough to produce magnetic order. Inrhodium, however, no ordering was observed in spite ofthe record-low and record-high temperatures of 280 pKand –750 pK.

Nevertheless, susceptibility measurements showedthat rhodium nuclei tend to order antiferromagneticallyat both positive and negative spin temperatures, whichcan be understood if the nearest-neighbor and next-nearest-neighbor exchange interactions are competing,as the NMR measurements suggest. It is then possiblethat the observed structure at T.0 is a type-I antiferro-magnet, as in copper and silver, while the ground stateat T,0 is an fcc antiferromagnet of the type II.

The spin-lattice relaxation of rhodium showed an in-teresting feature: t1 was clearly longer at T,0 than atT.0 (Hakonen, Vuorinen, and Martikainen, 1994).Such an effect had not been seen in similar measure-ments on silver. The observation was explained by thepresence of iron impurities in the sample, at the level of10 ppm. It was concluded that the large susceptibility ofRh nuclei modifies the scattering of conduction elec-trons by the impurities and leads to changes in t1.

F. Studies of other metals

Before beginning a thorough discussion of experimen-tal and theoretical results on copper, silver, andrhodium, we should like to mention some interestingrecent results on nuclear magnetism in other metals(see Sec. XII for a more detailed discussion). TheBayreuth group has made extensive measurements onthe properties of the metallic compound AuIn 2 at lowtemperatures in fields above the critical field for super-conductivity (Herrmannsdorfer and Pobell, 1995;Herrmannsdorfer, Smeibidl, Schroder-Smeibidl, and Po-bell, 1995). Data on nuclear specific heat, magnetic sus-ceptibility, and the measured NMR spectra gave strongevidence for a ferromagnetic transition at TC535 mK.The magnitude of TC was unexpectedly large, perhapsby an order of magnitude. The magnetically orderingspins in AuIn 2 are the In nuclei, which form a simplecubic lattice. Unlike Cu, Ag, and Rh, ordered nuclei ofAuIn 2 are in thermal equilibrium with the conductionelectrons.

Possible indications of a transition to an ordered statehave been reported in scandium (Suzuki et al., 1994;Koike et al., 1995). The quadrupolar interaction is im-portant in this metal. As a result, Sc is a good exampleof a three-dimensional Ising model in a hexagonal lat-tice, so that the spin Hamiltonian is qualitatively differ-

Page 10: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

10 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

ent from that in the noble metals. It seems clear thatnew data will be forthcoming, but it is too early to dis-cuss these systems in detail at present. Interesting NMRresults on highly polarized thallium nuclei have been ob-tained by Eska and Schuberth (1987) and Leib et al.(1995).

Several short reviews on nuclear ordering in metalshave been published earlier (Oja, 1987; Lounasmaa,1989; Hakonen, Lounasmaa, and Oja, 1991; Oja, 1991;Hakonen and Vuorinen, 1992; Hakonen, 1993; Ra-makrishnan and Chandra, 1993; Hakonen and Lounas-maa, 1994; Pobell, 1994; Steiner, Metz, Siemensmeyer,et al., 1996).

II. BASIC PRINCIPLES OF POSITIVE AND NEGATIVESPIN TEMPERATURES IN NUCLEAR MAGNETISM

A. Spin temperature

The idea of spin temperature originates from thework of Casimir and Du Pre (1938). Their work wasextended by Bloembergen and Wang (1954), who sug-gested that there may be two temperatures within thespin system, one, TZ , relating to the distribution of theZeeman energy, and the other, Tss , relating to the en-ergy of spin-spin interactions. When TZ5Tss it is mean-ingful to speak of a common spin temperature; other-wise a distinction between TZ and Tss should be made.A large number of experiments have verified theoreticalpredictions based on the concepts of TZ and Tss . Com-prehensive reviews have been written on spin tempera-ture and its implications (Goldman, 1970; Wolf, 1979;Abragam and Goldman, 1982). We also refer the readerto the paper by Van Vleck (1957), in which the basics ofpositive and negative spin temperatures are discussed ina clear and simple way.

1. Zeeman temperature

The meaning of the Zeeman temperature TZ , associ-ated with the Hamiltonian HZ52\gB•( iIi , can beillustrated using an energy-level diagram. Since the nu-clei are well localized in their lattice sites, Boltzmannstatistics is appropriate for a description of the spin sys-tem. The energy-level diagram for a spin I5 1

2 assemblyis shown in Fig. 5(a) in the case when the Zeeman en-ergy dominates over spin-spin interactions. Thenumber of nuclei in the different energy levelsEm52\gmB , m52I . . . 1I , is proportional toexp(2Em /kBTZ). At positive temperatures, the numberof nuclei in a higher level of energy is always smallerthan that in a lower energy level.

2. Temperature of the interaction reservoir

The temperature of the interaction reservoir Tss canbe understood in a similar way. Let us consider the situ-ation in high external fields, where spin-spin interactionsare but a small perturbation to the Zeeman Hamil-tonian. In the energy-level diagram, interactions spliteach Zeeman level into a number of sublevels. The char-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

acteristic width of the splitting is on the order of thespin-spin interaction energy, and the number of sublev-els is proportional to (2I11)N, which in practice is infi-nite (N is the number of spins).

Mathematically, the meaning of Tss as the tempera-ture of the interaction reservoir Hss8 is described by thedistribution function (i.e., the density matrix)

r5exp~2HZ /kBTZ!exp~2Hss8 /kBTss!. (7)

Thus the distribution of the nuclei in the sublevels ofeach Zeeman level is determined by the same Tss . Ingeneral, Tss can differ from TZ . Such a situation is illus-trated in Fig. 5(b). In thermal equilibrium, however,Tss5TZ , as in Figs. 5(a) and 5(c).

The detailed form of Hss in the density matrix is asubtle point. Hss8 is that part of the full spin-spin interac-tion Hss which conserves the Zeeman energy, i.e., com-mutes with HZ . When the Zeeman and interation reser-voirs are not in mutual thermal equilibrium, Hss8 must beused in the density matrix, as is written into Eq.(7). However, in thermal equilibrium Tss5TZ , andthe full Hss is the appropriate quantity; thenr5exp@2(HZ1Hss)/kBT# . Exchange forces of the formIi•Ij , such as the Ruderman-Kittel interaction [see Eq.(3)], commute with HZ , while the dipolar Hamiltoniandoes not. This has the important consequence that therewill be thermal contact, so-called thermal mixing, be-tween the Zeeman and interaction reservoirs.

B. Heat reservoirs, spin-lattice relaxation,and thermal mixing

The various parts of the spin Hamiltonian can bethought of as thermal reservoirs with a certain heat ca-pacity and temperature. Figure 6(a) schematically illus-trates the situation in which the Zeeman and interactionreservoirs are in mutual thermal equilibrium. This istypically the case for fields that are low or comparable tothe local field B loc . A single temperature then describesthe spin system. The lattice vibrations and conductionelectrons form a heat reservoir at the temperature Te .

Equalization of Te and TZ is obtained via a relaxationprocess,

d~TZ21!/dt5t1

21~Te212TZ

21! , (8)

where t1 is the spin-lattice relaxation time. In metals,t1 is obtained from

t15k/Te , (9)

where k is the Korringa constant (Korringa, 1950). Thisequation is valid for fields much higher than B loc and fornot too large values of B/Te . In low B , spin-lattice re-laxation becomes faster. The situation can still be de-scribed by Eq. (9) if a field dependency is incorporatedinto the Korringa constant through

k~B !

k`5

B21B loc2

B21aB loc2 . (10)

Page 11: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

11A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

FIG. 3. At right: Neutron-intensity contour diagram of copper for the Bragg reflections, as a function of time and the externalmagnetic field: (solid curves), (1 1

313); and (dashed curves) (1 0 0). The number of neutrons collected per second is marked on the

contours. From Annila et al. (1990). At left: Spin structures of copper for the [011 ] alignment of the magnetic field, as given byEqs. (159) and (140). (a) B=0: antiferromagnetic k15(p/a)(1,0,0) structure consisting of alternating ferromagnetic planes. (b)

0,B,Bc/3: structure with ordering vectors k15(p/a)(1,0,0) and 6k56(p/a)(0, 23 , 2

3 ), illustrated for B50.17Bc ; (0 23

23) and

(1 13

13 ) reflections are equivalent under fcc symmetry. (c) B5Bc/3: the up-up-down structure with 6k56(p/a)(0, 2

3, 23) order. (d)

High-field state with three ordering vectors (p/a)(1,0,0), (p/a)(0,1,0), and (p/a)(0,0,1). The spin structures, which are consistentwith the neutron-diffraction data, were taken from theoretical calculations by Viertio and Oja [1990a (structures a–c), 1987(structure d)].

According to measurements as well as theoretical calcu-lations, the constant a varies in the range a52–3 (Gold-man, 1970). Spin-lattice relaxation is also faster thanthat given by Eq. (9) at large B/Te (Jauho and Pirila,1970; Bacon et al., 1972; Shibata and Hamano, 1982).

In high fields, thermal contact between the Zeemanand interaction reservoirs is weak. It is therefore usuallymore appropriate to describe the spin system in terms ofthe block diagram illustrated in Fig. 6(b). As was men-tioned above, the interaction reservoir now consists ofthose interactions which commute with HZ .

The relaxation of HZ and Hss8 towards Te is deter-mined by different time constants, t1 and t1ss . Theformer is described by the usual high-field Korringa con-stant, Eq. (9). The relaxation time t1ss is not of practicalimportance for experiments discussed in this paper.

The direct thermal contact between HZ and Hss8 iscalled thermal mixing, and the corresponding time con-stant tm is the thermal mixing time. Thermal mixing is

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

caused by spin-spin interactions that do not conservethe Zeeman energy. Usually tm is determined by thedipolar interaction. Processes that contribute to thermalmixing are single spin flips (terms Ii

1Ijz and Ii

2Ijz) and

double spin flips (Ii1Ij

1 , Ii2Ij

2 ; see Sec. XI.A.2 for defi-nitions). An isotropic exchange interaction commuteswith HZ and therefore does not cause thermal mixing.

The constant tm depends strongly on the magneticfield, becoming exceedingly long at high B . The reasonfor this is easy to understand. The energy must be con-served during the relaxation process. If the change in theZeeman energy is large, it cannot be counterbalancedeasily by a corresponding but opposite change in theinteraction energy. Various theoretical models predict,approximately, tm't0exp(B2/b2) or tm't0exp(B/b),where the constant b is on the order of B loc and t0 is onthe order of the spin-spin relaxation time t2 (Goldman,1970). An experiment in which the field dependence of

Page 12: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

12 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

tm was measured for silver will be discussed in Sec.XI.A.

C. Demagnetization to the ordered state

We discuss next the process during which a sample isdemagnetized into the magnetically ordered state. Thenuclear-spin system is first polarized in a high field and a

FIG. 4. Right half: Phase diagram of nuclear spins in silver atT.0 in the magnetic field vs temperature plane. ‘‘Inside’’ thefull curve the spin system is antiferromagnetically ordered.The phase boundary is drawn on the basis of measurements byHakonen, Yin, and Nummila (1991), assuming that its shape isthe same as in the magnetic field vs entropy plane. The spinconfigurations at right represent a two-sublattice, type-I anti-ferromagnet in B50 (lower structure) and in a finite field(higher structure). These configurations, with the ordering vec-tor parallel to the external magnetic field, are consistent withthe neutron-diffraction data of Tuoriniemi, Nummila, et al.(1995) and with theoretical predictions in low fields but not inhigh fields (Viertio, 1992; Heinila and Oja, 1993a). The exter-nal magnetic field is pointing down. Note that the gyromag-netic ratio of silver nuclei is negative and therefore spins Ii

align antiparallel to Bext at T.0. Left half: Phase diagram atT,0. The solid curve shows the boundary between the ferro-magnetic domain state (inside) and the paramagnetic region.From Hakonen, Nummila, Vuorinen, and Lounasmaa (1992).One of the possible domain configurations predicted byViertio and Oja (1992) is also shown.

FIG. 5. Equilibrium distribution of nuclei in the various Zee-man levels: (a) at a positive spin temperature TZ for a spinI5

12 system. Splitting into sublevels is due to spin-spin interac-

tions. (b) Situation when Tss.0 and TZ,0. (c) Equilibrium ata negative Tss5TZ .

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

low temperature to decrease its entropy below the criti-cal value for magnetic ordering. After the equilibriumpolarization with TZ5Te has been reached, the sampleis demagnetized down to B50, or to a low field, to pro-duce spontaneous nuclear ordering. The demagnetiza-tion is performed in a time much shorter than t1 toavoid warmup of the spin system caused by spin-latticerelaxation.

Let us first neglect, for simplicity, the spin-lattice re-laxation. When we change from the high initial field to alow, so-called mixing field Bm , we see that the only ef-fect of the demagnetization is to decrease the Zeemantemperature, so that the ratio B/TZ remains constant.At B5Bm , the thermal mixing time tm becomes suffi-ciently fast compared to the rate of the field sweep thatthere will be thermal contact between the Zeeman andinteraction heat reservoirs. As a result, thermal equilib-rium with TZ5Tss is achieved. It is only now that theinteraction reservoir is cooled. Before the mixing pro-cess it remained at the initial precooling temperature.

Thermal mixing is necessarily a nonadiabatic process,since the cold Zeeman temperature has to cool down thehot interaction reservoir. The degree of adiabaticity ischaracterized by the factor

Bm2 /~Bm

2 1B loc2 !, (11)

which is derived in Sec. IV.F. The closer to one thisfactor, the more adiabatic will be the thermal mixing. Itis obvious that it would be desirable to perform mixingin the highest possible Bm . In practice, however, ther-mal mixing is too slow a process to occur in fields muchlarger than B5B loc .

On the other hand, nuclear ordering also takes placein fields on the order of B loc . One may wonder whetherthe non-occurrence of nuclear ordering after demagne-tization to a certain low field might simply be becausethermal mixing has not yet taken place. This does notseem to be the case. Measurements on silver, for ex-ample, show that Bm'10B loc50.3 mT while the criticalfield for nuclear ordering is Bc'3B loc50.1 mT. Thusthe final cooling necessary for nuclear ordering takesplace in a field three times larger than the critical field.

FIG. 6. Schematic illustration of the Zeeman (HZ), spin-spin-interaction (Hss), and conduction-electron (HL) heat reser-voirs. The various thermal relaxation times, t1, tm , and t1ss ,are indicated. (a) In low fields, HZ and Hss form a single heatreservoir, but (b) in high fields they are decoupled.

Page 13: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

13A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Theoretical estimates and actual measurements showthat the nonadiabaticity associated with thermal mixingis on a few-percent level and that it is not prohibitive fornuclear ordering. A more detailed explanation is givenin Sec. IV.F. The present discussion is, of course, anoversimplification in that thermal mixing is, in reality, agradual process, which proceeds in a range of fields dur-ing demagnetization. The steep field dependence oftm , however, makes the steplike process a good ap-proximation.

The form of the isentropes in the ordered state is animportant question. Mean-field calculations, to be dis-cussed in Sec. XV.B.5. for antiferromagnetic type-I spinstructures or more generally for the so-called permanentspin structures show that isentropes in the ordered stateare vertical in the magnetic field vs temperature plane,as shown in Fig. 7. The temperature T then stays con-stant during demagnetization in the ordered state. Fig-ure 7 also depicts the entropy increase due to thermalmixing at B5Bm .

D. Negative spin temperatures

The main ideas associated with negative spin tempera-tures can again be illustrated conveniently by an energy-level diagram. For the sake of simplicity, we consider adiagram for a noninteracting I5 1

2 spin system in an ex-ternal magnetic field. The same ideas can be appliedreadily to interacting spin assemblies in an arbitraryfield.

At the absolute zero, T510, all nuclei are in theground state with mW 5\gI parallel to B. As the tempera-ture is increased, keeping B constant (see Fig. 8), nucleibegin to flip into the upper energy state and, atT51` , there is an equal number of spins in both levels.If the energy is increased further, the reversed distribu-

FIG. 7. Demagnetization into the antiferromagnetically or-dered state along isentropes in the magnetic field vs tempera-ture plane. Thermal mixing takes place in the field B5Bm ,larger than the local field B loc . In silver, Bm'0.25 mT,Bc50.1 mT, B loc50.035 mT, and TN5560 pK.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

tion of nuclear spins can still be described by the Boltz-mann factor but now with T,0. Finally, when ap-proaching zero from the negative side, T→20,eventually only the highest energy level is populated.

As Fig. 8 shows, the transition from positive to nega-tive spin temperatures is smooth and takes place viaT56` . The positive and negative infinite temperaturescorrespond to the same spin arrangement.

Since heat has to be added to the spin system atT.0 to reach a state with T,0, negative temperaturesare actually hotter than positive ones. If two spin assem-blies, one at T.0 and the other at T,0, are broughtinto thermal contact, heat will flow from the latter to theformer.

1. Production of negative temperatures

Purcell and Pound (1951) first produced negative tem-peratures in the nuclear-spin system of LiH. Their ex-perimental procedure was, in fact, a simple one. If thedirection of the external magnetic field is reversed sofast that spins do not have time to redistribute amongtheir energy levels, the spin assembly enters a state thatcan be described by a negative temperature. After thefield flip, there are more nuclear moments antiparallel tothe field than parallel to it. In terms of the energy-leveldiagram, the spin populations at the Iz5 1

2 and Iz52 12

levels are exchanged.A sufficiently rapid field reversal is fast in comparison

with the spin-spin relaxation time t2, which is the ap-proximate length of time between successive spin flips inlow external fields. If the field change is slower, the spinsare able to follow adiabatically the field reversal, and

FIG. 8. Energy-level diagram of nuclear spins (I512) in silver

or rhodium at selected temperatures when B=constant. Spinpopulations in the Iz5

12 and Iz52

12 levels are shown at both

positive and negative absolute temperatures. Rapid reversal ofthe external magnetic field B interchanges the energy levelsand produces a negative spin temperature.

Page 14: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

14 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

negative temperatures will not result. Demagnetizationwould just be followed by remagnetization to the posi-tive starting temperature.

To estimate the rate of field change needed to pro-duce negative temperatures one may assume that onlythe low-field region B51B locz→2B locz has to bepassed in a time much shorter than t2. On the otherhand, t2 is approximately the Larmor period of the spinsin the local field B loc created by neighboring nuclei.Thus the critical field sweep rate is on the order of

DB

Dt'

gB loc2

2p. (12)

In fact, during the quick field flip the Boltzmann distri-bution of the spins breaks down and, for a short mo-ment, the spin system cannot be assigned a temperature.

Once the spins have been brought to the T,0 state,they remain there for a long time. The spin system re-laxes towards the positive lattice temperature through apassage via T=7`, at a rate determined by the relativelylong t1.

Only the Zeeman temperature is reversed in the fieldflip; the temperature of the interaction reservoir is not.See Fig. 5(b) for an illustration of such a situation. Anirreversible entropy increase results when HZ warms theinteraction reservoir to negative temperatures, produc-ing the equilibrium depicted in Fig. 5(c). This thermalmixing process is similar to that described earlier (seeSec. II.B). A more detailed discussion is presented inSec. IV.G.

Other methods for producing negative spin tempera-tures will be discussed in Secs. XIV.A and XIV.D.

2. Thermodynamics at T,0

Some of the fundamental aspects of negative tempera-tures have been described by Ramsey (1956). Our dis-cussion draws mostly from his text.

From the thermodynamic point of view, an essentialrequirement for the existence of a negative temperatureis that the entropy S not be a monotonically increasingfunction of the internal energy U . In fact, whenever(]S/]U)B,0, T51/(]S/]U)B,0 as well. At negativetemperatures an increase in U corresponds to a decreasein S , while the reverse is true when T.0 (see Fig. 9).For negative temperatures to occur, there must be anupper limit to all allowed energy states of the system,otherwise the Boltzmann factor exp(2Em /kBT) doesnot converge for T,0. Nuclear spins satisfy this require-ment, since there are 2I11 Zeeman energy levels. Inaddition, the elements of the assembly must, of course,be in thermodynamic equilibrium among themselves sothat the system can be described by the Boltzmann dis-tribution and thereby assigned a temperature. The ther-mal equilibrium time t2 among the nuclear spins them-selves must be short compared to the time t1 ofappreciable ‘‘leakage’’ of energy to or from other sys-tems. In silver, for example, t1514 h at Te5200 mKwhile t2510 ms.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Nuclear-spin assemblies are different from some morecommon systems for which the temperature in Kelvinsdescribes the average energy. Consider, for example, thekinetic energy E . When the temperature is raised to-wards infinity, E increases without an upper limit.Therefore the crystalline lattice or conduction electronscannot be brought to a temperature T,0: the energy ofthe system would be infinite.

At T,0 the second law of thermodynamics,T5DQ/DS , can just as well be used for thermometry(see Sec. II.E). In this case, DQ,0 when the entropyincreases and the system radiates energy at the Larmorfrequency of the nuclei, while the populations of the twoenergy levels tend to equalize.

At T510 and at T520 there is complete but oppo-site order, i.e., the entropy is zero in both cases. WhenT56` , S has its maximum value R ln2 for the spinI5 1

2. At T,0, adiabatic demagnetization heats the spinsystem instead of cooling it, as happens when T.0.Similarly, for experiments on polarized nuclei at T,0,the spin system must be heated to the hottest negativetemperature to achieve maximum polarization, while atpositive temperatures the spins must be cooled.

The difficulty of heating a ‘‘hot’’ system at negativetemperatures is analogous to the problems of cooling acold system at positive temperatures. An NMR absorp-tion experiment at T.0 becomes an NMR emission ex-periment at T,0. This effect was observed with the7Li nuclei by Purcell and Pound (1951), and it providedthe key proof that negative spin temperatures actuallyhad been produced in the nuclear-spin system: the emis-sion peaks at T,0 gradually became weaker until, atT56` , the emission and absorption canceled out be-cause of equal populations of the energy levels. Slowlyincreasing absorption spectra were then observed as thespin system ‘‘cooled’’ towards room temperature.

When two systems are brought into thermal contact,heat always flows from the hotter to the colder body.The order of temperatures on the absolute Kelvin scale,from the coldest to the hottest, is thus +0 K, . . . +300 K,. . . 6` K, . . . 2300 K, . . . 20 K. The system cannotbecome colder than 10 K since it is not capable of giv-

FIG. 9. Entropy S as a function of the internal energy U forthe two-level system of Fig. 8. Negative slope of the S/R ln2 vsU/NmB curve corresponds to a temperature T,0.

Page 15: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

15A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

ing up more of its energy; likewise, it cannot becomehotter than 20 K because the system is unable to absorbmore energy.

Near the absolute zero, 1/T or logT is sometimes usedas the temperature function but, when T,0, logT is notsuitable. However, on the inverse-negative scale whereb521/T , the coldest temperature, T510, correspondsto b52` , and the hottest temperature, T520, tob51` . On this scale the algebraic order of b and theorder from cold to hot are identical; the system passesfrom positive to negative Kelvin temperatures throughb520→10. The choice of the function, b521/T , en-sures that a colder temperature is always to the left of ahotter one. This inverse-negative scale thus runs in an‘‘orderly’’ fashion from the coldest to the hottest. Thethird law of thermodynamics emerges ‘‘naturally’’ by theimpossibility of reaching the positive or negative ends ofthe b axis.

The theorems and procedures of statistical mechanics,such as the use of the partition function or the quantum-mechanical density matrix, apply equally to systems atnegative temperatures. By examining the statisticaltheory by which the Boltzmann distribution is derived,we see that there is nothing objectionable a priori in theparameter 1/kBT being negative; T,0 simply meansthat the mean energy of the system is higher instead ofbeing lower than that corresponding to equal popula-tions among the energy levels at T56` .

Figure 10 illustrates the entropy S , the internal energyU , and the specific heat CB of a two-level spin assembly(see Fig. 8) as a function of b521/T in units ofkB /umuB , calculated in the usual way [see Eqs. (25a) and(26a)]. The entropy has its maximum value S5R ln2 atb560 because both energy levels are equally popu-lated (the spin polarization p50). The specific heatCB is zero at b52` and at b51` since all spins oc-

FIG. 10. Entropy (dotted curve), internal energy (full curve),and specific heat (dashed curve) plotted as a function of2umuB/kBT for a nuclear-spin system of two energy levels(I5

12, as for Ag and Rh, and umu5 1

2\ugu).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

cupy their lowest or highest energy level and no moreheat can be removed or absorbed, respectively. Atb560, CB50 as well because a very large change oc-curs in T for a very small change in the spin configura-tion. The internal energy has its minimum value atb52` (T510) and its maximum at b51`(T520).

Although systems at negative temperatures can betreated without difficulty by thermodynamics and statis-tical mechanics, albeit with certain modifications to theKelvin-Planck formulation of the second law of thermo-dynamics (Ramsey, 1956), the occurrence of systems atT,0 is infrequent because of the rather restrictive re-quirements for a thermodynamic assembly to be de-scribed by a negative temperature. Population inversionalone is not sufficient. A Boltzmann distribution of theparticles among the energy levels, i.e., internal thermalequilibrium, is needed before a temperature can be as-signed to the assembly. Lasers do not operate at nega-tive temperatures. Laser beams, however, have beenused to achieve very low temperatures, for example, 100nK in an assembly of sodium atoms (Kasevich and Chu,1992).

3. Nuclear ordering at T,0

At T510 an isolated nuclear-spin assembly has thelowest, and at T520 the highest, possible energy. Thisfeature can be put into a more general thermodynamicbasis. As the external magnetic field B is reducedtowards B loc and ultimately even to B50, thedipole-dipole and exchange forces gradually take over,and the spin order begins to change from that forced byB to an arrangement determined by mutual interactions.During this spontaneous adjustment of spins the entropyincreases, according to the general principles governingthermodynamic equilibrium, until S reaches amaximum, while the magnetic enthalpy H5U2BMstays constant.2 One must thus consider the variation ofentropy under the restriction of a constant enthalpy,i.e., seek an extremal value of S1bH where b is aLagrange multiplier. By differentation, dS1bdH=0, so that b52dS/dH521/T . Therefore, one findsS2H/T52G/T for the thermodynamic potentialreaching an extremum; G is the Gibbs free energy.When T.0, G5H2uTuS , and the extremum is a mini-mum, since S assumes its maximal value at equilibrium.When T,0, G5H1uTuS , and the Gibbs free energyobviously reaches a maximum.

The tendency to maximize rather than to minimizethe energy is the basic difference between negative andpositive temperatures. This produces a profound effecton the spin structure into which a system spontaneouslyorders below the transition temperature when T.0 or

2Enthalpy, which is the sum of the internal and magnetic en-ergies, is constant since the system is isolated. In terms of themicroscopic Hamiltonian, Eq. (1), this would mean^H&5constant.

Page 16: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

16 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

‘‘above’’ it when T,0. In silver, for example, thenearest-neighbor antiferromagnetic Ruderman-Kittelexchange interaction favors antiparallel alignment of thenuclear magnetic moments and thus produces antiferro-magnetism when T.0. At T,0, since the energy nowmust be maximized, the very same interactions causeferromagnetic nuclear order.

E. Thermometry with nuclear spins

Measurements of the temperature of a nuclear-spinsystem are based on the second law of thermodynamics,T5DQ/DS . This relation can be used equally well atpositive and negative temperatures. A practical way toadminister a small heat pulse DQ on the spin system isto expose it to an alternating magnetic field of a suitablefrequency f for a time Dt . Then DQ } B1

2x9(f)Dt , whereB1 is the amplitude of the alternating field. The crucialdifference between the behavior at T.0 and that atT,0 is that in the former case spins absorb energy fromthe alternating field whereas in the latter case spins emitenergy to this field. Hence at T,0, DQ,0 when theentropy increases. This is borne out also by the fact thatx9(f).0 at positive temperatures while the opposite istrue for negative temperatures. The common feature atT.0 and at T,0 is that, when exposed to an alternat-ing magnetic field of the resonant frequency, the popu-lations of the two energy levels tend to equalize.

Measurements of DS ensuing from the heat pulse areusually made by relating S to the spin polarization pbefore and after the pulse; p can be determined fromthe area of the NMR peak.

Thermal conductivity within the nuclear-spin systemin a metal is such that, over atomic distances, equilib-rium is reached quickly. However, on a macroscopicscale thermal diffusion is very slow. This means that inany experiment all parts of the sample must be subjectedto similar treatment so that different regions of thespecimen will be at the same spin temperature.

A thorough discussion of nuclear-spin thermometryon copper, silver, and rhodium is given in Sec. IV.

III. EXPERIMENTAL TECHNIQUESFOR SUSCEPTIBILITY MEASUREMENTS

A. Principle of brute force nuclear cooling

The basic principle of nuclear cooling is the same asthat for paramagnetic salts (de Klerk, 1956), but thereare very significant differences in practice. Becausenuclear magnetic moments are about 2000 times smallerthan their electronic counterparts, it is more difficult toproduce significant changes in the entropy of thenuclear-spin system by external means. For example, toreduce the entropy by 5%, one must start from an initialtemperature Ti510 mK and an initial magnetic fieldBi56 T, if copper is used as the working substance; forparamagnetic salts the corresponding values are 1 K and1 T. Fortunately, the starting conditions for nuclear

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

cooling today can be reached rather easily with dilutionrefrigerators and superconducting magnets (Lounasmaa,1974).

The main advantage of nuclear cooling is, of course,the very low temperature that can be reached. Nucleialign spontaneously, owing to their mutual interactions,well below 1 mK. Because spontaneous ordering is thelimit of any cooling process, temperatures in the submi-crokelvin region can be reached by nuclear-demagnetization. The current record, 280 pK, was ob-tained by demagnetizing a rhodium specimen in acascade nuclear-demagnetization process (Hakonen,Vuorinen, and Martikainen, 1993). For cerium magne-sium nitrate, the weakest paramagnetic salt, the orderingtemperature is slightly below 2 mK.

The technique of nuclear cooling consists of first mag-netizing the sample isothermally from zero to Bi>5 T ata low initial temperature Ti<20 mK. The mixing cham-ber of the precooling dilution refrigerator and thenuclear stage are then thermally isolated from eachother by means of a superconducting heat switch, andthe magnetic field is reduced to a low final value Bfadiabatically. The nuclear sample is thereby cooled to

Tf5~Ti /Bi!~Bf21B loc

2 !1/2 , (13)

where B loc is 0.36 mT for copper, 35 mT for silver, and34 mT for rhodium and represents the effective interac-tions between the nuclei. The exact definition of B loc willbe given by Eq. (29). Thus, for example, if one startsfrom Bi58 T and Ti516 mK and demagnetizes to Bf50, the final temperature is 0.7 mK using copper nuclei.

The basic equations of brute force nuclear cooling willbe discussed here only briefly; for more detail we referthe reader to other publications (Huiskamp and Lounas-maa, 1973; Lounasmaa, 1974; Betts, 1976; Andres andLounasmaa, 1982; Pickett, 1988; Pobell, 1992a, 1992b),which also describe dilution refrigeration and other rel-evant cryogenic techniques in detail.

In an external magnetic field B , the 2I11 equi-distant nuclear energy levels are given by Em52mNgNBm52\gBm , where mN55.05310227 Am 2

is the nuclear magneton, gN5m/I is the nuclear g factor,and m runs from 2I to +I , with I denoting the nuclearspin. The partition function of the system is

Z5F(m

exp~2Em /kBT !GnNA

, (14)

where n is the number of moles of the sample, kB isBoltzmann’s constant, and NA is Avogadro’s number;nNA is thus the number of magnetic nuclei in the speci-men. The population of the mth energy level is given by

P~m !5nNAexp~2Em /kBT !/ (m

exp~2Em /kBT ! . (15)

In the approximation Em!kBT , the entropyS5kB](T lnZ)/]T becomes

S5nR ln~2I11 !2nLB2/2T2 , (16)

Page 17: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

17A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

where the nuclear Curie constant per moleL5NAI(I11)mN

2 gN2 /3kB . In a constant field, the

nuclear heat capacity CB5T(]S/]T)B is then, in thesame approximation, given by

CB5nLB2/T2 . (17)

Equation (14) shows that Z and therefore alsoP(m), S , and CB are functions of B/T only for all valuesof B and T . During adiabatic demagnetization(DS50) from Bi and Ti to Bf , S remains constant, andtherefore Bi /Ti5Bf /Tf ; the final temperature is thengiven by

Tf5TiBf /Bi . (18)

If demagnetization is carried out to a low or zero field,Bf must be replaced in this equation by (Bf

21B loc2 )1/2;

we then obtain Eq. (13). Similarly, if Eq. (17) is to beused in low fields, B must be replaced by (B21B loc

2 )1/2.After demagnetization, the nuclear-spin system begins

to warm up owing to the unavoidable external heat leakdQ/dt5Q . By observing that T5Q/CB , one can calcu-late the time Dt during which the nuclear spins warmfrom Tf to a higher temperature T . The result is

Dt5~nLBf2/Q !~Tf

212T21! , (19)

where Eq. (17) has been employed. The relation showsthat 1/T is a linear function of time, provided that Q isconstant. It should be noted that Dt is proportional toBf

2 ; demagnetization should not be carried out to fieldslower than is necessary to reach the desired temperatureTf .

The rate at which equilibrium is established betweennuclear spins and conduction electrons is governed bythe spin-lattice relaxation time t1, defined by Eq. (8).For most metals t1 is on the order of seconds at 10 mK;for insulators t1 is days or even weeks. It is thus clearthat metals must be used for brute force nuclear refrig-eration. The lattice heat capacity, proportional to T3, istotally negligible below 10 mK. One may therefore as-sume that Te represents the common conduction-electron and lattice temperature.

The short spin-lattice relaxation time in metals is dueto conduction electrons that act as intermediaries be-tween the nuclear spins and the lattice. Only electronsnear the Fermi surface contribute; their number is pro-portional to Te making t1 proportional to 1/Te . Thisresults in Eq. (9), t15k/Te (Korringa, 1950). As wasdiscussed in Sec. II.B, the Korringa behavior is modifiedin fields comparable to B loc and for high B/Te . In prac-tice, among the elemental metals, only copper, indium(Symko, 1969), silver, thallium (Angerer and Eska, 1984;Eska and Schuberth, 1987), scandium (Suzuki et al.,1994; Koike et al., 1995), and rhodium have been usedfor brute force nuclear cooling.

The unavoidable external heat leak Q into aconduction-electron system has an important effect onthe equilibrium between Te and T . If Q is large, thespin-lattice relaxation process is not sufficiently rapid

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

for cooling the conduction electrons adequately, andTe2T will be large. One can derive the equation

~Te /T !215kQ/nL~Bf21B loc

2 ! (20)

starting from the equations governing spin-lattice relax-ation, Eqs. (8) and (9) (Lounasmaa, 1974).

To obtain significant refrigeration of conduction elec-trons, one should not carry out nuclear demagnetizationall the way to Bf50 but should stop at some intermedi-ate field value. In fact, the lowest Te is reached by de-magnetizing to

Bf~opt!5~kQ/nL!1/2 ; (21)

in this case Te /T52. In the derivation of this result ithas been assumed that Bf@B loc .

The very lowest temperatures have been obtained bymeans of cascade nuclear magnetic cooling techniques.In cryostats of this type, the first nuclear stage acts as aprecooler for the much smaller second stage. Three cas-cade refrigerators, namely, those at Helsinki (Ehnholm,Ekstrom, Jacquinot et al., 1979; Ehnholm et al., 1980;Huiku et al., 1986), at Risø (Jyrkkio et al., 1989), and atthe Hahn-Meitner-Institut in Berlin (Tuariniemi, Num-mila, et al., 1995), were employed in the experiments tobe described in detail in this review. Other machines ofthe cascade type have been built in Julich (Mueller et al.,1980), Tokyo (Ishimoto et al., 1984), and Bayreuth(Gloos et al., 1988); in these cryostats the aim is usuallyto cool conduction electrons or 3He. For useful reviews,see Huiskamp and Lounasmaa (1973); Pickett (1988);and Pobell (1988).

B. Helsinki cryostat

The main parts of the Helsinki cryostat are schemati-cally illustrated in Fig. 11 (Ehnholm, Ekstrom, Jacquinotet al., 1979; Ehnholm et al., 1980). The apparatus con-sists of a dilution refrigerator and two nuclear stages,operating in series. During its active life of nearly twodecades, the cryostat has naturally undergone manymodifications (Huiku et al., 1986; Hakonen and Yin,1991; Oja, Annila, and Takano, 1991).

The dilution unit in this apparatus was designed tobe used as the precooling stage of a nuclear-demagnetization cryostat. The main objectives in itsconstruction were therefore a rigid structure and a highcooling power down to about 10 mK. The home-maderefrigerator was able to reach 7 mK without an externalheat load and to cool the nuclear stage in an 8-T fielddown to 10 mK. Between the mixing chamber and thefirst nuclear stage there is a superconducting heat switchmade of a piece of bulk tin.

In the first design of the apparatus (Ehnholm, Ek-strom, Jacquinot et al., 1979; Ehnholm et al., 1980), theupper nuclear stage was assembled from 10 moles ofcopper wire, 0.5 mm in diameter and insulated with fi-berglass. The early version of the second nuclear stage,i.e., the sample, was made of 2000 copper wires, 0.04 mm

Page 18: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

18 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

FIG. 11. Cascade nuclear-demagnetization cryostat in Helsinki. The main parts below the mixing chamber of the dilution refrig-erator, inside the vacuum jacket, are shown in more detail at right. The second nuclear stage is also the sample; it is connected tothe first stage by welding. From Ehnholm et al. (1980) and Huiku et al. (1986).

in diameter and insulated by oxidation; the measuredresidual resistivity ratio RRR=200.3 This wire was cho-sen because its spin-lattice relaxation time had previ-ously been measured (Aalto et al., 1972). The weight ofthe sample was 2 g. At their upper ends, the thin copperwires were welded together and connected to the firstnuclear stage by means of copper foils, without any heatswitch. Therefore, thermal contact between the conduc-tion electrons in the two nuclear stages was good.

The electronic temperature Te in the sample was de-termined by the temperature of the large upper nuclearstage, by the thermal conductivity between the twonuclear stages, and by the heat leak Q to the secondstage [see Eq. (20)]. Te5250 mK was reached, corre-sponding to Q50.3 nW to the sample. The heat input tothe first nuclear stage was 3 nW.

The magnetic fields for operating the nuclear stageswere generated by two superconducting solenoids. The

3RRR=r(300 K)/r(4.2 K).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

upper stage could be magnetized to a maximum field of8 T. The magnet had field-compensated volumes at bothends; the upper low-field region was at the location ofthe heat switch and the mixing chamber, and the lowerlow-field region was at the site of the second nuclearstage. The effective amount of copper in the 8-T fieldwas neff52.5 mol; this early version of the first stage wasthus relatively small compared with other nuclear de-magnetization cryostats built in Helsinki (Gylling, 1971;Veuro, 1978) or elsewhere. In later versions of the ap-paratus, neff was increased.

The superconducting magnet surrounding the lowernuclear stage produced a maximum field of 7.3 T. It alsohad two field-compensated regions: the upper one at thejoint between the two nuclear stages, so that a heatswitch could have been assembled there if desired, andthe lower one starting 110 mm below the center of themagnet, at the site of the SQUID that was used for mea-surements of the nuclear susceptibility. Inside the 7.3-Tsolenoid, there is a mu-metal tube to reduce the effect ofthe field trapped in the superconducting magnet afterthe current has been removed.

Page 19: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

19A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

The difficulty in the first experiments (Ehnholm et al.,1980) was the short spin-lattice relaxation time, 10–20min, which caused the system to warm up rather rapidlyand apparently prevented the observation of nuclear or-dering. To remedy this problem, the upper wire bundlewas replaced in 1981 with a solid copper rod, 22322mm2 cross section and 30 cm long, designed by P. Rou-beau (Huiku et al., 1986). To reduce eddy-current heat-ing during demagnetization, 0.6-mm wide slits, spaced1.6 mm apart, were cut on each side of the square rod(see Fig. 11). The mass of the new nuclear stage was 1.1kg, which corresponded to an effective size of 10 mol ofcopper in the operating magnetic field of 8 T. The ad-vantages of a bulk copper nuclear stage are that (l) it isrigid, (2) it is relatively easy to make, (3) it is a goodconductor of heat, (4) all kinds of thermal contacts canbe made to it, (5) it is devoid of insulating materialsneeded in a wire bundle to eliminate eddy currents dur-ing demagnetization, and (6) it is easy to handle.

Two essential improvements were observed (Huikuet al., 1986) after replacing the old ‘‘wire bundle’’ firstnuclear stage by the bulk copper rod: the conduction-electron temperature was lowered from 250 to 50 mK,and the minimum heat leak to the first nuclear stage wasreduced from 2 to 0.5 nW. A drawback of the bulknuclear stage was an enhanced time-dependent heatleak: the low 0.5 nW level was reached after the cryostathad been kept below 4 K for 6 weeks; two days aftercooldown the heat leak was still 10 nW. Time-dependentheat leaks have been discussed by several authors(Loponen et al., 1981; Zimmermann and Weber, 1981;Pobell, 1982, 1992a; Schwark et al., 1983; Kolac et al.,1985).

Conduction-electron temperatures were measured bymonitoring the susceptibility of platinum wires usingpulsed NMR techniques (Lounasmaa, 1974; Pobell,1992a). The design of the thermometer (Huiku et al.,1986; Hakonen and Yin, 1991) was similar to that usedby the Julich group (Buchal, Hanssen et al., 1978). Com-mercial NMR electronics were employed (PLM-4, RV-Electronics, Veromiehentie 14, 01510 Vantaa, Finland).The static field for the thermometer was produced by asmall, unshielded, and end-compensated solenoid madeof multifilamentary NbTi wire. The Pt signal was cali-brated against the superconducting transition tempera-tures of Be and W samples in a fixed-point device(Soulen and Dove, 1979). All thermometers were in-stalled on top of the first nuclear stage.

In experiments on silver and rhodium, reaching suffi-cient initial polarization of the sample was a much moredifficult task than it was with copper. This is due to thesmallness of the magnetic moments of silver andrhodium nuclei as well as to their long spin-lattice relax-ation times. To solve these problems, an even more mas-sive version of the first nuclear stage was made, withneff521 mol (Oja, Annila, and Takano, 1991), and therigidity and heat conductivity of the thermal link be-tween the two nuclear stages was improved (Hakonenand Yin, 1991).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

C. Experimental procedure

In studies of nuclear magnetism of copper the experi-mental procedure, illustrated in Fig. 12, was as follows(Huiku et al., 1986). After the superconducting heatswitch had been turned on,4 the first nuclear stage wasmagnetized to 8 T. The dilution refrigerator then pre-cooled, in about 40 h, the neff510 mol copper stage toabout 10–12 mK, after which the heat switch was turnedoff to isolate the nuclear stages from the dilution unit.

Next, the second nuclear stage was magnetized, in 40min, to 7.3 T and, starting simultaneously, the first stagewas demagnetized to Bf50.1 T in about 5 h, with arapidly decreasing dB/dt towards the end of demagne-tization. This procedure was possible because the heat ofmagnetization of the small second stage could be ab-sorbed easily by the first nuclear stage at 0.1 T. In prin-ciple, it would have been preferable to do the precoolingwith the second-stage field on, but the magnet could notoperate in the persistent mode. This resulted in a largeboiloff from the 4He bath when the magnet was oper-ated at the full current.

After waiting for 1.5 h at the 0.1-T field, demagneti-zation of the first stage was continued to Bf520 mT,producing a temperature Te550 –100 mK in theconduction-electron system of the sample. The secondnuclear stage was then demagnetized rather rapidly. Atypical time from 7 T to zero field was 20 min; this keptpolarization losses in the sample low, 0.5–1.5% duringthe field sweep.

After demagnetization of the lower magnet, theconduction-electron temperature of the copper speci-men remained constant for several hours because of thehigh cooling capacity of the large first nuclear stage. Fur-thermore, the low heat resistance (3–6 K 2/W) in thethermal link between the specimen and the first stageand the small external heat leak (<30 pW) to the sampleguaranteed a constant Te along the whole secondnuclear stage and thus a constant relaxation time t1 ofthe nuclei while they warmed up.

The residual trapped field in the 7.3-T solenoid sur-rounding the sample was a serious problem in the veryfirst measurements. A too high field would keep thesample above Bc even at Bext50. In addition, the slowrelaxation of the trapped field increased noise in themeasuring system. In tests it was found that immediatelyafter demagnetization the residual field could be as highas 1 mT, decreasing only slowly, in about 20 min, to60.1 mT. Outside the center of the magnet even stron-ger fields were measured. The problem was overcome bysurrounding the specimen with a mu-metal tube whosepermeability, after heat treatment, was 40,000 at room

4This means that a current was allowed to flow in the smallsuperconducting solenoid surrounding the switch; the magneticfield so produced forced tin into the normal state, where itconducts heat well. When the current was switched off, thestrip of tin returned to its superconducting state, at which itconducts heat very poorly; the switch was then in its off state.

Page 20: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

20 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

FIG. 12. Schematic illustration, on a temperature vs entropy diagram, of the cascade nuclear-demagnetization procedure forcooling an assembly of silver or rhodium nuclei to positive or negative nanokelvin temperatures. (A→ B) Both nuclear stages arecooled to Ti515 mK by the dilution refrigerator and, simultaneously, the first stage is polarized in a strong magnetic field Bi

58 T. (B→ C) The nuclei of the first stage, made of 20 mol of copper, are adiabatically demagnetized to Bf5100 mT, whichproduces a low temperature Tf5Ti(Bf /Bi)'200 mK. Towards the end of demagnetization, the second nuclear stage, that is, thesilver or rhodium sample, is magnetized to 8 T. (B→ D) The 2-gram specimen of thin polycrystalline foils cools in the field B58 T, by thermal conduction, to Tf'200 mK. (D→ E) The specimen is demagnetized from 8 T to 400 mT, whereby the spins coolinto the low nanokelvin range @T'(400 mT/ 8 T) 200 mK=10 nK], thermally isolated by the slow spin-lattice relaxation (t1514 h) from the conduction electrons, which are anchored to 200 mK by the first nuclear stage at C. When demagnetization of thespecimen was then continued to zero field (not shown in the illustration), the record temperature of 280 pK was reached inrhodium. In silver, dipole-dipole and exchange interactions produced antiferromagnetic order at the Neel temperature TN5560pK. (E→ F) A negative temperature can be produced in the spin system of silver or rhodium nuclei by reversing, instead ofdemagnetizing to zero, the 400-mT magnetic field in about 1 ms. The rapid field flip causes some loss of polarization (that is,increase of entropy). When demagnetization was continued to zero field, the record temperature of 2750 pK was reached inrhodium (not shown in the diagram). In silver, dipole-dipole and exchange interactions produce ferromagnetic order at the Curietemperature TC521.9 nK. (F→ G→A) The system starts to lose its negative polarization, crossing in a few hours, via infinity,from negative to positive temperatures. (C→A) The first nuclear stage warms slowly, under the Bf5100 mT field, fromTf5200 mK towards 15 mK. A new experimental sequence can then be started.

temperature. In high magnetic fields, the mu-metal satu-rates and has no effect, whereas in low fields, even im-mediately after demagnetization, it shields the sampleeffectively against external fields less than 8 mT. Withthe tube in place, the residual field decreased below 0.05mT in 1 min, and in 2 min it was below 0.02 mT. Formeasurements on silver and rhodium, in which the criti-cal field is lower than in copper, an additional magneticshield was installed on the outer surface of the vacuumcan. This decreased the remanent field to 2–5 mT.

After demagnetization of the lower magnet, thesample was in a field of 10 mT, generated by a smallsuperconducting solenoid inside the mu-metal tube. Await of about 1 min allowed the noise in the SQUID

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

system to decrease sufficiently for final demagnetizationto the low field, usually between 0 and 0.3 mT, in whichnuclear ordering was to be investigated. The spin systemwas thereby cooled below 50 nK in copper, while theconduction electrons remained at 50 mK.

In experiments on silver and rhodium, polarizing thesample was slower than with copper. The cooling proce-dure was therefore modified. After the first nuclearstage had been demagnetized down to a low field, typi-cally 65 mT, which produced Te5100–150 mK, thesample was polarized in the 7.3-T field for about 20 h.Polarizations up to 94% were obtained in silver (Ha-konen and Yin, 1991). Even longer times were neededto polarize the rhodium sample. The first-stage demag-

Page 21: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

21A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

netization was then continued down to 35–45 mT whilethe sample was kept in 8.3 T, produced by a new second-stage magnet, for 30 h. Polarization of 83% was therebyreached in rhodium (Vuorinen, Hakonen, Yao, andLounasmaa, 1995). Longer polarizing times were pre-vented by the need to refill the 4He bath of the dewar.Demagnetization of the silver and rhodium samplesdown to zero field produced spin temperatures well be-low 1 nK.

D. Sample preparation

The samples used for studies of nuclear ordering bythe Helsinki group have usually been polycrystalline. Aspecimen made of several thin foils or wires gives astronger NMR signal than a bulk single crystal becausethe electromagnetic field is screened by the skin effect inbulk material. In addition, eddy-current heating, whenthe sample is being demagnetized, is a less severe prob-lem in thin foils and wires. The specimen should obvi-ously be as pure as possible so that impurities will notaffect the ordering process. It is difficult, however, toknow a priori what is the purity needed, and in practicethe best available samples were always used. The purityaffects the low-temperature thermal and electrical con-ductivity, which are proportional to each other accord-ing to the Wiedemann-Franz law. In spite of increasededdy-current heating, high thermal conductivity is desir-able so that the heat of magnetization, created when thesample is polarized in a high field and low temperature,can efficiently be conducted away. In practice, however,this is not very critical in reasonably pure (99.99+%)copper and silver, nor in 99.96+% pure rhodium owingto its relatively small nuclear magnetic moment. Thesample was usually made of thin foils rather than wires,since good thermal contacts are easier to achieve withfoils. A further important aspect of impurities is theireffect on the spin-lattice relaxation time.

1. Quenching of magnetic impurities by internal oxidation

During the early years of the Helsinki investigationson copper (Ehnholm, Ekstrom, Jacquinot et al., 1979;Ehnholm et al., 1980) it was observed, by measuring thespin-lattice relaxation time t1, that below 2 mT magneticimpurities speed up the relaxation process. The high-field relaxation time t1(15 mT) was longer, by a factorbetween 4 and 7, than the experimentally importantzero-field relaxation time t1(0). This effect, togetherwith the fact that the conduction-electron temperaturewas as high as 250 mK, presumably prevented definiteobservation of nuclear ordering.

Several approaches were tried to increase t1(0). Thefinal success grew out of the dismaying observation(Huiku, Loponen et al., 1984) of an anomalously rapidspin-lattice relaxation in a very pure copper specimen.An impurity analysis, using atomic absorption tech-niques, showed that this material had less than 1.3 ppmof magnetic contaminants: Mn<0.1, Cr=0.3, Fe=0.8,and Ni<0.1 ppm. The ratio r5t1(Bhigh)/t1(0), withBhigh515 mT@B loc(Cu)=0.36 mT, at first was as large as

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

100 but, after the sample had been prepared by oxidiz-ing it at 950 °C for 3 h under a pressure of 0.1 mbar ofdry air, the spin-lattice relaxation rate became muchslower in low fields; the new ratio r52.6 was essentiallythe same as the calculated value for a spin system withno impurities (Goldman, 1970).

Therefore the effect of magnetic impurities, which aidin the relaxation process at low fields, can be quenchedby a suitable heat treatment that changes the magneticimpurities into their nonmagnetic oxides; Fig. 13 illus-trates the spin-lattice relaxation time for the new speci-men as a function of the external magnetic field, afternormal annealing and after internal oxidation. The ef-fect is dramatic and was of crucial importance for ob-serving nuclear ordering in copper.

After having observed the effect of internal quenchingon the low-field spin-lattice relaxation time, the Helsinkigroup routinely applied the oxidation technique to allsamples studied for nuclear ordering. It is likely that theeffect has its origin in the same processes that typicallyled to improved low-temperature electrical conductivityin oxygen-annealed copper, as had been found previ-ously by many authors (Fickett, 1974). Measurements onseveral samples carried out in Helsinki over a long timesuggest that there can be a correlation between theRRR and the low-field spin-lattice relaxation ratet1(B low)21, but the effect is not clear from Fig. 14.

It has been emphasized (Fickett, 1974) that oxygenannealing lowers the chemical purity although its effectscan usually be described by an improvement of electricalpurity. Recent studies (Shigematsu et al., 1992) on ultra-pure copper (99.99999%) show, indeed, that oxygen an-nealing can be very harmful to the low-temperatureelectrical conductivity when the copper matrix is ex-tremely pure. By analogy, one may speculate that suchheat treatment could shorten t1 at low fields in an ultra-pure metal.

Oxygen annealing has also been observed to increasethe RRR of silver (Ehrlich, 1974). Anticipating that an-nealing would increase t1 in low fields, the Helsinki

FIG. 13. Spin-lattice relaxation time of a high-purity coppersample: n, before selective oxidation; s , after selective oxida-tion. From Huiku et al. (1986).

Page 22: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

22 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

group selectively oxidized all silver samples prepared fornuclear ordering experiments. A typical treatment took18 h at 750 °C in an oxygen atmosphere of 0.2 mbar. Thetemperature during the heat treatment was somewhatlower than for copper because of the higher vapor pres-sure of silver. The annealing resulted in RRR’s up to 900for 25-mm thick foils and up to 1500 for 125-mm thickfoils, both of 99.99+% nominal purity (Hakonen andYin, 1991; Oja, Annila, and Takano, 1991). Measure-ments of the spin-lattice relaxation rate of the 25-mmthick foils showed that r5t@B@B loc(Ag)=35 mT]/t1(0)was within the expected value between 2 and 3 for purematerial.

In rhodium, effects due to impurities could not befully removed by internal oxidation. The annealing pro-cedure for 25-mm thick foils was 16 h at 750 °C in anoxygen atmosphere of 0.4 mbar (Hakonen, Vuorinen,and Martikainen, 1993, 1994). The first sample con-tained about 100 ppm of iron, and its Korringa constantdecreased from k510 sK to 0.06 sK in small magneticfields, i.e., r5170. The second sample had a nominalpurity of 99.96+% and the total quantity of magneticimpurities was less than 15 ppm. It had r550, which wassufficient for carrying out extensive zero-field suscepti-bility measurements. The RRR’s were 250 for the firstsample and 530 for the second. From these values theeffective iron contents were estimated as 14 ppm and 6ppm, respectively.

2. Thermal contact

A problem closely connected with sample preparationis joining the specimen to the thermal link between thesample and the first nuclear stage. Several different tech-niques have been used for this purpose: tungsten-tipinert-gas (TIG) welding (Ehnholm et al., 1980), diffusionwelding (Huiku et al., 1986), and electron-beam welding

FIG. 14. Ratio of the spin-lattice relaxation times measured inhigh and low fields, r5t1(Bhigh)/t1(B low), for samples of cop-per, silver, and rhodium with different residual resistivity ratios(RRR). Data are from Huiku, Loponen et al. (1984) and fromJyrkkio et al. (1989) for Cu, from Hakonen and Yin (1991) andOja, Annila, and Takano (1991) for Ag, and from Hakonen,Vuorinen, and Martikainen (1994) for Rh.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

(Hakonen and Yin, 1991; Oja, Annila, and Takano,1991). Conventional TIG welding produces low contactresistances between similar metals after annealing(Muething et al., 1977). TIG welding is, however, unsuit-able for joining foils and dissimilar metals owing to alloyformation.

Diffusion welding is particularly convenient with foils.To join two pieces of copper, for example, a preliminaryjoint is first made at a low temperature (400 °C) bypressing the foils together with a stainless-steel clamp.Once the pieces stick together, the clamp is removedand the final diffusion welding is done at a high tem-perature (1000 °C), perhaps simultaneously with oxygenannealing. Diffusion welding of dissimilar materials ispossible, for example, between copper and aluminum(Bunkov, 1989), but the procedure can be difficult if themelting temperatures Tm of the two metals are very dif-ferent. One can then coat the material of higher Tm witha thin layer of the metal having the lower melting tem-perature. Thermal joints between silver and rhodiumfoils have been made this way (Vuorinen, Hakonen,Yao, and Lounasmaa, 1995) by first sputtering a 0.2mm layer of silver on rhodium and then making the dif-fusion weld between the silver foil and the sputteredsurface of the rhodium foil.

Electron-beam welding is practical if the foils to bejoined are not too thin; together the two pieces shouldbe at least 1 mm thick. Good thermal joints betweencopper and silver are easily achieved; a contact resis-tance of 0.55 mV mm 2 has been measured for a jointbetween these metals (Yin and Hakonen, 1991). To re-lease welding stresses, the joint was annealed for 3 h at750 °C; the contact resistance then decreased to 0.2mV mm 2. The thickness of the alloyed layer at the jointwas studied by an electron microscope using back-scattering, and the layer was found to be 1–30 mm thick,depending on the position. The layer is thus rather thinin comparison with the typical spot size of 0.5 mm of theelectron beam, making it possible to obtain a low con-tact resistance.

The actual setup for nuclear ordering experiments onsilver, depicted in Fig. 15, illustrates many of the aspectsdiscussed above. The sample consists of 48 silver foils 25mm thick. The foils were folded into U shapes andgrouped together with other foils as shown, in order toincrease the rigidity of the structure. The foils were elec-trically insulated from each other by sprinkling themwith a 7-mm layer of SiO 2 powder. The sample was con-nected to a thermal link made of silver by a diffusionweld, which was further improved by electron-beamwelding. The link was machined into the shape of across, again to increase the mechanical rigidity of thewhole structure. This is important for reducing vibra-tional heat leakage in high magnetic fields, as well as fordecreasing the noise during susceptibility measurements.The link was connected to the first nuclear stage byelectron-beam welding.

We have discussed sample preparation in great detailbecause it was of crucial importance for the success ofthe experiments.

Page 23: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

23A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

E. SQUID measurements of the NMR signaland the low-frequency magnetic susceptibility

Apart from the neutron-diffraction data on copperand silver (see Secs. VII and IX), all experiments onnuclear ordering in copper, silver, and rhodium arebased on measurements of the dynamic magnetic sus-ceptibility x(f)5x8(f)2ix9(f), where x8(f) is the dis-persion at frequency f and x9(f) is the absorption.5 Inhigh fields, x9(f) and x8(f) are usually referred to as theNMR absorption and dispersion. In low fields, when thewidth of the x9(f) signal becomes comparable to thefrequency at the maximum of x9(f), description ofx9(f) as the nuclear-magnetic-resonance absorptiondoes not correspond to the conventional picture, and itis better to call x9(f) the absorptive part of the dynamicsusceptibility. It is both fortunate and important thatmeasurements of the dynamic susceptibility alone giveinformation not only on the magnetic behavior but alsoon the thermodynamics of the nuclear-spin system,which is isolated sufficiently well from the conductionelectrons.

The NMR measurements on copper, silver, andrhodium were made by using a superconducting quan-tum interference device (SQUID). In this type of study,SQUID NMR has several advantages over conventional

5Here x(f) refers to a diagonal component of the matrixx(f), which, in most cases, is perpendicular to the externalmagnetic field.

FIG. 15. Schematic view of the silver sample and its connec-tion to the copper nuclear-demagnetization stage. The NMRcoils are located over the lower end of the sample. From Ha-konen and Yin (1991).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

NMR techniques based on a field-effect-transistor(FET) preamplifier. SQUIDs offer greater sensitivity atlow frequencies because a SQUID magnetometer mea-sures directly the change in the magnetic flux caused bychanges in magnetization, whereas the conventionalmethod responds to dM/dt , where M is the magnetiza-tion. In particular, for NMR on metals at low tempera-tures, working at low frequencies is advantageous fromthe point of view of penetration depth and eddy-currentheating of the sample by the rf field. SQUID NMR isideal for studying systems that have long spin-lattice re-laxation times and short spin-spin relaxation times, i.e.,fast dephasing times. The bandwidth of a system em-ploying a SQUID can be made very large. As an ex-ample, in the apparatus of Chamberlin et al. (1979), thesame instrument could be used both for NMR down to10 kHz and for low-field electron paramagnetic reso-nance up to 1 GHz. Meredith et al. (1973) and Webb(1977) have discussed the basic principles of SQUIDNMR and have presented typical applications. For amore recent review on the use of SQUIDs in low-frequency applications, see Ryhanen et al. (1989).

Ehnholm et al. (1980) introduced and developed theSQUID NMR technique for investigations of nuclear or-dering on copper. The methods employed in later mea-surements of Cu, Ag, and Rh were variants of this origi-nal scheme. The measuring system consists essentially ofa pickup coil connected to an rf SQUID, which works asa sensitive preamplifier over a large frequency range.The dynamic susceptibility is recorded by varying theexcitation frequency rather than the external field. Dur-ing some measurements in high fields, however, corre-sponding to Larmor frequencies close to 200 kHz, theconventional field-sweep method was employed.

A typical coil assembly for measuring the magneticresponse of the sample is shown in Fig. 16 (Oja, Annila,and Takano, 1991), which illustrates two systems, onelongitudinal, the other transverse, for detecting different

FIG. 16. Coils used in the NMR measurements in (a) horizon-tal and (b) vertical external fields. The dashed vertical linesindicate sections to which (a) the excitation and (b) the staticfield coils extend. From Oja, Annila, and Takano (1991).

Page 24: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

24 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

components of the susceptibility. The longitudinal-coilarrangement, Fig. 16(a), is used for measuring the NMRsignal in horizontal static fields, created by a saddle coilin the xy plane, whereas the transverse system, Fig.16(b), is employed when the static field is vertical.

In the longitudinal measuring geometry, the verticalpickup coil consists of two 4-mm high, seven-turn sole-noids, which were wound in opposite directions, with an11-mm separation to reduce direct feedthrough of theexcitation signal. The inductance of the pickup coil wasmatched to the 2-mH input inductance of the SQUID.The excitation field was generated by a small solenoidoutside. For calibrating the SQUID response, a smallsolenoid was wound on the lower half of the pickup coilbelow the sample.

In the horizontal geometry, Fig. 16(b), the transverse,saddle-shaped 30-mm-long figure-of-eight pickup coilwas made of four loops. Two transverse excitation coilswere wound, one coaxial with the pickup coil and theother at a right angle to it. A transverse saddle-shapedcalibration coil was also wound on the lower half of thepickup loop.

The crossed-coil system can be employed only in non-zero fields, where the macroscopic magnetization rotatesaround Biz. In zero field, when there is no longer anypreferred axis, the response of the spins vibrates in thedirection of the excitation and thus no signal is detectedif the excitation and pickup coils are crossed. Thereforethe zero-field measurements were made with a coaxialcoil system. Another possibility would be to use a rotat-ing excitation, generated by feeding into the two crossedexcitation coils currents with 690° phase shift. The re-sponse to positive and negative frequencies could thenbe measured separately (Ehnholm et al., 1980).

All coils in the setup of Fig. 16 were wound on a cy-lindrical support of 10 mm diameter, made of Stycast1266. The coil system was mounted inside the mixing-chamber radiation shield surrounding the sample regionso that the support did not make any contact with thespecimen. In earlier experiments on copper, the measur-ing coils were wound on a cylindrical coil former madeof silver and thermally anchored to the first nuclearstage. The advantage of this scheme is a better fillingfactor. This design was abandoned, however, in the mea-surements on silver and rhodium, since a clearance be-tween the sample and the coil former guaranteed that noheat could leak to the specimen, which has a very smallthermal capacity.

The pickup coils were usually wound of normal-metalwire, such as insulated silver or copper, rather than su-perconducting material. The advantage of a resistivepickup wire is that the input circuit of the SQUID,which is shown schematically in Fig. 17, then forms ahigh-pass filter and noise below the cutoff frequency isattenuated, so that it is easier to maintain the opera-tional position of the SQUID; the quality of the signal isimproved as well. In practice, by a suitable choice ofpickup wire, the cutoff frequency fL5Rp /(2pLs) canbe chosen between 10 Hz and 1 kHz with a typicalLs52 mH inductance of the signal coil in the SQUID.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

10 Hz is slightly below the important frequency rangefor NMR on silver and rhodium, while fL'1 kHz is suit-able for copper. The upper cutoff frequency of the de-tection system could be adjusted by a small shunt resis-tor across the signal coil of the SQUID.

Pickup loops made of superconducting wire were triedas well, but with variable degrees of success. In somesetups flux jumps in the superconducting pickup wirewere suspected of introducing much noise, although inother cases a superconducting pickup coil made of 70-mm multifilamentary NbTi wire in a CuNi matrixworked without problems (Huiku and Loponen, 1982).A superconducting pickup loop would also allow a di-rect measurement of the dc magnetization of the sample.This possibility was tried in the very first studies on cop-per, but the noise was several orders of magnitude largerthan the expected signal.

The dynamic susceptibility x(f) was measured usingfrequency sweeps across the resonance in a constant ex-ternal magnetic field. The block diagram of the measur-ing system is shown in Fig. 18. Detection of the NMRspectra was controlled by a computer. When measuringa spectrum, an oscillator was set to sweep through thefrequency range of interest. The ac output was fed intothe excitation and compensation coils and to the refer-

FIG. 17. Low-temperature parts of a SQUID-NMR measuringsystem. Lp and Ls are the inductances of the pickup and signalcoils, respectively. R is the resistance of the pickup coil. A 3-V resistor protects the signal coil of the SQUID from high-frequency disturbances. Components forming an LC resonantcircuit, the so-called tank circuit, are shown on the right.

FIG. 18. Computer-controlled SQUID-NMR measuring sys-tem. The experimental procedure is explained in the text.

Page 25: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

25A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

ence input of the biphase lock-in analyzer. The in- andout-of-phase signals of the SQUID electronics weremeasured by a lock-in amplifier. The two phases weremonitored by a digital voltmeter and a recorder.

When the measuring frequencies were low, theSQUID was used in the flux-locked mode.6 This tech-nique made it possible to employ higher excitation lev-els, over half a flux quantum, which improved the signal-to-noise ratio. During measurements in the 100-kHzfrequency range the SQUID was not operated in thefeedback mode, but the voltage across the tank circuitwas monitored directly (Ehnholm, Ekstrom, Loponenet al., 1979; Ekstrom et al., 1979).

As the stability of the SQUID response during mea-surements is important, possible changes in the gain orphase of the detection system were checked after everyrecorded NMR spectrum by feeding a small current intothe calibration coil below the sample. This coil was alsoused to measure the overall frequency dependence ofthe gain and the phase. In a metallic sample an extraphase shift is caused by eddy-current shielding (Chap-man et al., 1957).

In practice, a limitation on the accuracy of the mea-surements was not intrinsic to the SQUID but due toother sources of noise. This is not surprising consideringthe fact that the pickup coil was supposed to measuresmall magnetic signals at a location that had been in an8-T field half an hour earlier. The noise level usuallyincreased permanently after the 8-T solenoid had beenenergized for the first time after cooldown to 4.2 K. Sev-eral sources of degraded performance were identified inthe various setups: (1) The noise in the measuring sys-tem, as well as the lowest temperatures achieved, weresensitive to the level of external vibrations. (2) Afterdemagnetization, the mu-metal tube around the samplerelaxed slowly and produced noise during the first fewminutes. (3) The remanent field of the 8-T sample mag-net drifted for several hours after the end of demagne-tization. This effect, and the remanent field itself, couldbe reduced by feeding a small opposite current to the8-T solenoid after demagnetization. In addition, thisprocedure accomplished some degaussing in the mu-metal shield. (4) When a superconducting pickup wirewas used, movements of trapped flux caused noise forseveral hours after demagnetization. (5) In some setups,the SQUID sensors were operated at 0.7 K, althoughtheir performance had been optimized for use at 4 K;this caused a small increase of noise. (6) In many cases,transfer of liquid helium to the main dewar warmed theSQUID to the normal state owing to its poor thermalanchoring to 4.2 K. When the SQUID returned to itssuperconducting state it trapped magnetic flux, gener-

6In the flux-locked mode, the flux penetrating the SQUIDring is kept constant. This is achieved by compensating for anychanges in the external flux by the use of a feedback circuitand a coupling coil. There is a linear relationship between thefeedback current (or voltage) and the external flux. For moredetails, see Lounasmaa (1974).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

ated by the field of the first-stage magnet, and the noiselevel of the sensor became very large. The problem wassolved by heating the SQUID to the normal state justbefore the final demagnetization of the sample. Duringthis phase of the experiment the first-stage field was solow that it did not disturb the SQUID. (7) In most set-ups the SQUID was mounted close to the top plate ofthe vacuum can so that the pickup wires had to passthrough the first-stage field, which had to be low toavoid increasing the noise in the measuring system.

IV. MEASUREMENT OF NUCLEAR-SPIN TEMPERATURE

To compare theoretical calculations with experimen-tal data, one must know the temperature of the spinsystem. When the nuclei are not in equilibrium with theconduction electrons, one cannot measure T by using anexternal thermometer. The thermal isolation of nuclearspins makes it possible, however, to obtain the spin tem-perature directly from the second law of thermodynam-ics, viz.,

T5DQ/DS . (22)

The experimental problem is then reduced to giving aheat pulse DQ of known magnitude and measuring theresulting change in entropy DS . A heat pulse to nuclearspins can be administered through NMR absorption.When an alternating magnetic field B1(t)5B1sin(2pft)is applied for a time Dt , the energy absorbed by thespins is

DQ5pfB12x9~f !Dt/m0 , (23)

where x9(f) is the absorptive part of the complex sus-ceptibility. To measure DS one needs to know the en-tropy before and after the pulse. Entropy can be deter-mined by using adiabatic sweeps between the fieldwhere the temperature is measured, which is usuallyzero, and a high field. Here this means a field consider-ably stronger than B loc so that one can use the knownequations for noninteracting spins. The details of the ac-tual procedures used were somewhat different in thevarious measurements on copper, silver, and rhodium.Our discussion in the following emphasizes the more re-cent measurements on silver and rhodium.

A. Calibration of polarization and entropy

The measurement of the polarization p is based onthe fact that in a sufficiently high field, B5Bcalib@B loc ,p is proportional to the integrated area of the NMRabsorption signal,

p5AE0

`

x9~f !df , (24)

where A is a calibration constant. In practice, Bcalib was1.0 mT in Cu (Huiku et al., 1986), 0.19 mT in Ag (Ha-konen and Yin, 1991), and 0.40 mT in Rh (Hakonen,Vuorinen, and Martikainen, 1993).

The constant A can be determined by first producingan equilibrium polarization in a field on the order of 7 T

Page 26: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

26 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

at a relatively high electronic temperature around Te51 mK. Te was measured using the pulsed NMR tech-nique on platinum wires (Lounasmaa, 1974; Pobell,1992a, 1992b). Around Te51 mK, it is possible to equi-libriate nuclear spins with the conduction electrons andto avoid thermal gradients between the sample and thePt-NMR thermometer attached to the first nuclearstage. A commercial Pt-NMR electronics unit was em-ployed (PLM-4, RV-Electronics, Veromiehentie 14,01510 Vantaa, Finland). The platinum thermometer wascalibrated against the superconducting transitions oftungsten and beryllium at Tc515.760.1 mK and 22.66 0.1 mK, respectively. The calibrated W and Besamples were supplied by Dr. R. J. Soulen of the U.S.National Bureau of Standards. The fixed-point devicewas similar to that described by Soulen and Dove(1979).

Polarization of the nuclear-spin system in the 7-T fieldcan be calculated from

I5 12 : p5tanhu ~Ag, Rh!, (25a)

I5 32 : p5 1

3 ~4 coth4u2cothu ! ~Cu!, (25b)

where u5g\B/2kBT . These equations are valid fornoninteracting spins, since interactions can be ignored inthe 7-T field. As soon as the polarization is known, theentropy can be calculated from the equations

I5 12 : S/R5ln22 1

2 @~12p !ln~12p !1~11p !ln~11p !# ,(26a)

I5 32 : S/R5u~cothu24 coth4u !1ln~sinh4u/sinhu ! .

(26b)

For I5 32 one has to eliminate u from the coupled equa-

tions (25b) and (26b).Finally, the sample is adiabatically demagnetized from

the field B57 T to the field B5Bcalib used for monitor-ing p . However, a correction must be made for the lossof polarization due to spin-lattice relaxation during de-magnetization. This can be done by watching the decayof *x9(f)df , as shown in Fig. 19, and then extrapolatingthe signal back to the midpoint of demagnetization. Thisgives the area corresponding to the initial polarizationwith good accuracy.

B. Calibration of susceptibility

To calculate the magnitude of the heat pulse [see Eq.(23)], one has to know x9(f) in absolute units. Calibra-tion of the polarization alone yields only the productAx9(f). To find x9(f) one can use the Kramers-Kronigrelation

x8~0 !5~2/p!E0

`

~x9~f !/f !df , (27)

where x8(0) is the measured static susceptibility. Herewe assume the usual situation in which the fieldB5Bcalib is applied along the z direction and the mag-netic response @x8(0) or x9(f)] is monitored using acoaxial measuring system in the y direction. Both y andz are further assumed to be along the principal axes of

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

the demagnetizing tensor of the sample. The signal isthen related to the polarization through the equation

x8~0 !5m0pMsat /@B1m0~Dy2Dz!pMsat# , (28)

where Msat is the saturation magnetization and m0 is thepermeability of free space. Dy and Dz are the demagne-tizing factors of the sample in the y and z directions,respectively; they can be estimated from the dimensionsof the specimen. The term containing Dy2Dz typicallyaffects the calibration of x8(0) by only a few percent.

C. Local field

The local field B loc is an important quantity in nuclearthermometry. This is highlighted by Eq. (13), which di-rectly yields the spin temperature after adiabatic demag-netization to zero field if the starting conditions andB loc are known. B loc describes the average strength ofthe local fluctuating field that is felt by a nucleus. There-fore, it determines the entropy reduction at high tem-peratures in the disordered phase [see Eq. (35)].

The local field is defined by

B loc2 /B25Tr$Hss

2 %/Tr$HZ2 %, (29)

where Hss describes the spin-spin interactions HZis the Zeeman term, and the trace Tr$ . . . % is taken overspin states. Assuming dipolar and isotropic interactions,where Hss5HD1HRK , given by Eqs. (2) and (3), onefinds

FIG. 19. Decay of *x9(f)df after demagnetization of the silvernuclear spins to B5Bcalib had ended at t50. In this experi-ment Bcalib50.77 mT. The equilibrium value of the polariza-tion in B57 T and at Te'1 mK can be obtained by extrapo-lating the signal to the midpoint of demagnetization(t52td). The inset in the lower left corner shows a typicalNMR signal consisting of 107Ag and 109Ag resonances. Thesignal was integrated from f1 to f2 to obtain p . The inset in theupper right corner displays the data points that, together withthe origin, fixed the polarization scale. From Oja, Annila, andTakano (1991).

Page 27: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

27A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

B loc2 5B loc,D

2 1B loc,RK2 , (30a)

where

B loc,D5m0

4p\gF I~I11 !(

jr ij

26G 1/2

, (30b)

B loc,RK5F 12 I~I11 !(

jJ ij

2 G 1/2

/\g . (30c)

The dipolar sums ( jr ij26 have been tabulated for cubic

lattices (see p. 226 in Goldman, 1970).There is an intimate connection between B loc and the

width of the NMR absorption line. Roughly speaking,both quantities are determined by the randomly fluctu-ating local field. Details depend, however, on the par-ticular mixture of magnetic isotopes present in the sys-tem (Van Vleck, 1948). Various NMR measurements athigh temperatures, above T51.5 K, have yielded B loc inseveral materials, as reviewed by Oja (1987).

B loc can also be deduced by measuring the field de-pendence of the adiabatic susceptibility. According totheory (Anderson, 1962, and references therein), thereis the relationship

xL'xT /~11B2/B loc2 ! (31)

between the longitudinal xL and transverse xT adiabaticsusceptibilities; possible modifications caused by zero-frequency absorption have been neglected (Anderson,1962; Huiku and Soini, 1983). If this result is applied inthe high-T limit, one may use the Curie law xT5C/Tand Eq. (13) to obtain the field dependence of the tem-perature. One then finds xL}(11B2/B loc

2 )23/2. Figure20 illustrates the measured xL vs B for silver. A fit to thetheoretical curve yields B loc544 mT. However, the valueB loc535 mT, inferred from two different NMR measure-ments (Poitrenaud and Winter, 1964; Oja, Annila, and

FIG. 20. Reduced longitudinal susceptibility x0 of silver vs Bat p50.50 [see Eq. (34) and its discussion]. The solid linedisplays a fit to the theoretical curve xL}(11B2/B loc

2 )23/2,which yields B loc544 mT. The small shift of the symmetry axisfrom zero gives 1.4 mT for the remanent field. Modified fromHakonen and Yin (1991).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Takano, 1990), is somewhat lower. The latter result wasadopted for the temperature measurements on silver(Hakonen and Yin, 1991).

There are several other possible ways to determinethe local field. For example, measurement of loss in po-larization caused by a rapid field reversal yields B locthrough Eq. (41).

Data on xL in different fields can also be used to de-termine the magnitude of the remanent field, as is shownin Fig. 20.

D. Consistency checks using high-T expansions

Equations (25a)–(26b) are valid when the Zeeman en-ergy is much larger than the interaction energy. In lowfields one has to use other formulas. Exact thermody-namical relations can be derived within the high-T ap-proximation. Although they are not valid in the mostinteresting range of high p , these approximations areuseful for checking the consistency of thermometry inthe low-p limit. Consult Van Vleck (1937) for additionalterms in high-T expansions.

The two leading high-T terms for p can be written as7

p5~B/m0Msat!x0 /@12~R1L2D !x0# , (32)

where

x05C/T (33)

is the zero-field susceptibility of the noninteracting sys-tem and C5L/Vm is the nuclear Curie constant [see Eq.(16)] per mole divided by the molar volume Vm . HereL and D are due to dipolar interactions: L5 1

3 is theLorentz constant, and D is the demagnetizing factor inthe direction of the external field. R is the strength ofthe RK force, defined in Eq. (6). The values of C andm0Msat have been collected into Table II for Cu, Ag, andRh.

The longitudinal susceptibility xL5m0(dM/dB) isobtained by calculating the derivative of Eq. (32):

xL5x0 /@12~R1L2D !x0# . (34)

The equation can also be inverted to give x0 in terms ofthe measured xL . This is a way to ‘‘reduce’’ the suscep-tibility of the interacting spin assembly to that of a vir-tual system with no interactions.

7The coefficient @12(R1L2D)x0#21 reduces to11(R1L2D)x0 in the high-T expansion. The two expres-sions are identical, however, to the order considered here.

TABLE II. Curie constant C513I(I11)m0\2g2r/kB and satu-

ration magnetization m0Msat for natural isotopic mixtures ofcopper, silver, and rhodium.

C (nK) m0Msat (mT)

Copper 562 1200Silver 2.0 45Rhodium 1.3 41

Page 28: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

28 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

The leading higher-order effects of the spin-spin inter-actions on the entropy S are given by

S5nR ln~2I11 !2nL~B21B loc2 !/2T2 . (35)

The spin temperature at B50, for example, is deter-mined by measuring the entropy or polarization in ahigh field, B5Bcalib@B loc . The relation betweenT(B50) and, say, p(B5Bcalib) can be obtained withinthe high-T approximation, Eqs. (32) and (35). One finds

T5CB/@m0pMsat~11B2/B loc2 !1/2# , (36)

showing the T } p21 dependence at small polarizations.In temperature measurements on copper, silver, and

rhodium, Eq. (36) was not employed to check the con-sistency of the data, but rather to fit the amplitude B1 ofthe alternating field used to give a heat pulse DQ to thenuclear spins, as described by Eq. (23) (Huiku et al.,1986; Hakonen and Yin, 1991; Vuorinen, Hakonen, Yao,and Lounasmaa, 1995). The value calculated from thecoil geometry could not be used because the mu-metaltube (see Fig. 11) used to shield the sample against theremanent field was found to give a frequency-dependentchange in the B vs electric-current relation of the heat-ing coil. A direct measurement of B1 was difficult tomake as well. Therefore, in practice, B1 was calibratedby forcing the data to follow the high-T expansion, i.e.,Eq. (36).

A useful check to ascertain the consistency of nuclearthermometry would be to measure T in a high field,B5Bcalib , by using the heat-pulse method. The resultcould then be compared with the temperature obtainedfrom Eq. (25a), valid for noninteracting spins. Alterna-tively, it would be possible to calibrate the excitationfield B1 in this same way. The physically more interest-ing behavior at low fields could then be measured with-out fixing the high-T behavior. The problem with thisscheme might be, however, that applying DQ would re-quire frequencies so high that the rf field would not pen-etrate the sample. Another possible source of problemscould be the frequency dependence of the B vs currentrelation of the heating coil observed in some experimen-tal setups.

E. Secondary thermometers

A direct measurement of the nuclear-spin tempera-ture is a somewhat complicated and time-consumingprocedure. If a secondary thermometer is available itoften provides a more convenient method.

As an early example, in the study of the field-orientation dependence of the static susceptibility ofcopper (Huiku and Soini, 1983), the previously mea-sured (Ehnholm et al., 1980) transverse susceptibilitywas employed as a secondary thermometer to establishthe T dependence of the longitudinal susceptibility.

During studies on silver, secondary thermometry wasused extensively (Hakonen and Yin, 1991; Hakonen,Nummila, Vuorinen, and Lounasmaa, 1992). A draw-back in the direct temperature measurement was that arather large heat pulse DQ was needed to obtain suffi-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

ciently accurate values of DS . In terms of p , the changewas typically 20% per one measurement of T , and there-fore the direct procedure was not suitable for determin-ing the ordering temperatures accurately.

Secondary thermometry in the silver experiments wasbased on the observation that the relationship betweenthe zero-field temperature T and the inverse polariza-tion determined at B5Bcalib is linear over a wide range,as shown in Fig. 21. The linear relationship is guaranteedby Eq. (36) at high temperatures, but almost the sametendency continues to lower temperatures. In addition,the low-T limit is known approximately: neglectingquantum fluctuations, p→1 when T→0. Based on ex-perimental data, a piecewise linear dependence wasused: 1/p2150.55T below T510 nK and 1/p50.65Tabove 10 nK, where T is expressed in nK’s. The samefunctional dependence was observed at negative tem-peratures as well, with T being replaced by uTu, althoughthe scatter of the data was larger.

F. Nonadiabaticities

Nuclear-spin thermometry assumes that the state ofthe system develops adiabatically when, for example, theexternal field is reduced. There are, however, sources ofnonadiabaticity, i.e., entropy increases, some of whichcan be avoided and some of which cannot.

Increased spin-lattice relaxation during a field ramp,due to eddy-current heating proportional to (dB/dt)2, isan obvious candidate for raising Te . The field changeshould be slow enough.

Other possible sources of nonadiabaticity include con-sequences from interactions with electronic magneticimpurities, spurious nuclear quadrupolar interactionscaused by lattice defects, and surface phenomena possi-bly associated with quadrupolar effects in small particles(Komori et al., 1986) or thin foils. These interactionsmay depend on the magnetic field and in some cases on

FIG. 21. Inverse nuclear polarization 1/upu of silver at B50.19 mT, plotted against the absolute value of temperatureat B50; s , T.0; d T,0. The dashed curve was obtainedfrom the high-T expansion [see Eq. (36)], which has beenjoined smoothly to the linear low-T part passing through p51. From Hakonen and Yin (1991).

Page 29: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

29A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

the rate at which B is being varied, for example, whencross relaxation takes place between the quadrupolarenergy levels that cross each other as the field is beingchanged.

Even in an ideal sample there is some nonadiabaticityowing to thermal mixing, which was described in Secs.II.B and II.C. In a typical nuclear-ordering experiment,demagnetization of the specimen is fast in comparison tot1, and the cold Zeeman reservoir cools the hot interac-tion reservoir only in a field comparable to B loc .

To make our discussion quantitative, we employ thehigh-temperature expansion although it is not valid un-der the actual experimental conditions with the spinsalmost fully polarized. However, the high-T approxima-tion should be sufficient to describe, for example, mea-surements during which the polarization scale is cali-brated.

After the nuclei in the sample have been polarized inthe 7-T field at submillikelvin temperatures, the systemcan be described by the density matrix

s512bH , (37)

where b51/kBT and H5HZ1Hss8 1Hss9. Here Hss8 is, as

before, the secular part of the spin-spin interactions, i.e.,the part that commutes with HZ (@Hss8 ,HZ#50), whileHss9 is the remaining, nonsecular part for which@Hss9 ,HZ# Þ 0. After demagnetization has started, thechanging field modifies the density matrix. Since dB/dtis fast in comparison to t1, which in turn is much fasterthan tm at B57 T, the heat reservoirs in Fig. 6 no longerare in thermal equilibrium. Furthermore, the order inthe off-diagonal parts of s , corresponding to the non-secular interaction Hss9 , is destroyed, and s acquires theform (Goldman, 1970)

s512aHZ2b iHss8 , (38)

where b i is the inverse temperature at the beginning ofdemagnetization when B5Bi and a5b iBi/B (withoutany B loc corrections). Therefore only the Zeeman sys-tem is cooled during demagnetization in high fields,while the interaction reservoir remains at the precoolingtemperature.

When B becomes comparable to B loc , the Zeemanlevels, broadened by the spin-spin interactions (see Fig.5), start to overlap and thermal mixing takes place be-tweenHZ andHss8 , which means that s assumes again itsequilibrium form given by Eq. (37). The equilibriumtemperature b f

21 will be somewhat higher than a21 be-cause the Zeeman system has to cool the interactionreservoir.

To calculate b f , we assume that thermal mixing takesplace at the constant field B5Bm . Although B de-creases continuously during demagnetization, this as-sumption is reasonable since the field dependence oftm is very steep. During thermal mixing the energy re-mains constant, viz.,

Tr$sH%5Tr$~12aHZ2b iHss8 !H%5Tr$~12b fH!H% .(39)

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Ignoring the small term b iHss8 , and recalling thata5b iBi /B , we find

b f5b iBiBm /~Bm2 1B loc

2 !

5@b iBi /~Bm2 1B loc

2 !1/2#3@Bm /~Bm2 1B loc

2 !1/2# .

(40)

The last equation should be compared with Eq. (13),valid for adiabatic demagnetization. We see that in thepresent case there is a nonadiabaticity factorBm /(Bm

2 1B loc2 )1/2, which results from thermal mixing

when B5Bm .In a two-isotope system, such as a natural sample of

copper or silver, the behavior is more complicated. Dur-ing demagnetization it is then meaningful to describe thespin assembly in terms of three heat reservoirs. Thesewill be discussed in Sec. XI.A.2 when describing themeasurements of the thermal mixing time in silver. Al-though the behavior of the two-isotope system is morecomplicated than that with only one isotope, the formerappears to be more adiabatic.

If the spin system is remagnetized after thermal mix-ing has occurred in a low field, the heat contact is againdisconnected when B.Bm . During further field in-creases, the interaction reservoir remains at the low tem-perature it had when B5Bm , while HZ warms up inproportion to B . If a second demagnetization is thenperformed, the Zeeman system finds a cold interactionreservoir at B5Bm and there is only a small irreversibil-ity. There is some entropy increase, however, even inthis case, as the off-diagonal order in s , correspondingto the nonsecular Hss9 , is lost during field cycling. Onealso has to remember that the interaction reservoirwarms towards the lattice temperature with the relax-ation time t1ss , which is 2–3 times shorter than t1 in anideal specimen (Goldman, 1970) and possibly even lessin a real sample containing impurities.

Apart from direct measurements of tm in silver (Oja,Annila, and Takano, 1990), effects due to thermal mix-ing have been discussed (Ehnholm et al., 1980; Soini,1982) only briefly in the context of nuclear ordering ex-periments in metals. In insulators this process has beenstudied extensively (see the references in Abragam,1961). Let us now calculate the entropy increase due tothermal mixing in silver. We make our estimates for asingle-isotope system; the nonadiabaticity for a two-isotope assembly should be smaller. To define the ther-mal mixing field Bm , we choose a field at whichtm(B)51 s. Measurements (Oja, Annila, and Takano,1990) yield Bm'0.25 mT. The nonadiabaticity factorBm /(Bm

2 1B loc2 )1/2 then becomes 0.99. Thermal mixing

thus leads to a nonadiabatic 1% decrease in polarizationat B5Bm [see Eq. (32)] and a 2% decrease in the en-tropy reduction [see Eq. (35)]. In addition, there is anadiabatic 1% decrease in p at B5Bm50.25 mT, causedby the conventional (Bm

2 1B loc2 )1/2 term in Eq. (35). We

conclude that, at least in the low-polarization limit,nonadiabaticity due to thermal mixing in silver leads to arather small irreversibility, which, however, is significantfor accurate measurements.

Page 30: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

30 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

In the first set of experiments on nuclear-ordering incopper (Ehnholm et al., 1980), a nonadiabatic 7% maxi-mum loss in polarization was observed when a highlyordered spin system was demagnetized well below 1 mT,close to B loc50.36 mT. The effect was attributed tothermal mixing. This large loss was found only for thefirst demagnetization. After remagnetization to 1 mTand demagnetization to zero field for the second time,the loss was much smaller (Soini, 1982). Nonadiabatici-ties were observed in later measurements on copper aswell (Huiku et al., 1986). This time, however, polariza-tion losses were attributed to the ordering process ratherthan thermal mixing, and field sweeps between B51 mTand zero field in the disordered state, at entropies above0.61R ln4, were found to be adiabatic. Neutron-diffraction experiments (Annila et al., 1992) have clearlyshown that, indeed, during the transition from the(1 0 0) phase to the (0 2

323) phase in fields around B

50.11 mT, a large nonadiabaticity occurs. Thereshould, however, be some polarization loss caused bythermal mixing, too, although it may occur in fieldsabove B51 mT.

G. Production of negative spin temperaturesby a rapid field reversal

Although a rapid field reversal was used in the famousexperiment by Purcell and Pound (1951, see Sec. II.D.1)in which negative temperatures were first produced, itwas only rather recently that Oja, Annila, and Takano(1991) demonstrated that this method is feasible forstudying nuclear magnetism in metals. Hakonen et al.(Hakonen, Yin, and Lounasmaa, 1990; Hakonen, Num-mila, Vuorinen, and Lounasmaa, 1992; Hakonen andVuorinen, 1992) did a lot of work to improve this tech-nique for observing nuclear ordering in silver at T,0.For example, a special coil assembly and radiationshields were constructed to reduce eddy-current shield-ing.

The field-reversal procedure was somewhat compli-cated owing to the fact that the NMR spectra were mea-sured in a transverse field Bx5200 mT, created by asmall saddle-shaped coil, while the field inversion wasperformed using a solenoid producing a longitudinalfield Bz5400 mT. First, before the field flip, the initialpolarization was determined from the NMR spectrummeasured in the transverse 200-mT field (see Fig. 22).Next, the longitudinal solenoid was remagnetized to Bz5400 mT in 40–60 s, and Bx was simultaneously re-duced to zero. Bz was then changed to 2400 mTquickly, in 1 ms. The field reversal increased the internalenergy of the spin system considerably; the energy dur-ing this process is absorbed from the external magneticfield. Finally, Bz was slowly reduced to zero and Bx wassimultaneously increased to its original value of 200mT. The measured NMR emission spectrum (see Fig.22) then showed that, indeed, the spin temperature andthe nuclear polarization were negative. Final demagne-tization was then performed by reducing Bx to zero.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

The rapid reversal of the 400-mT field always resultedin some loss of polarization in the nuclear-spin system.This is illustrated in Fig. 23. The inversion efficiency wasabout 95% at small polarizations around pi50.4 but de-creased to 80% or lower at pi50.8. Therefore studies ofsilver at T,0 were limited to negative polarizations upto p5 –0.65.

The observed polarization loss is explained, at leastpartly, by thermal-mixing effects, which were discussed

FIG. 22. NMR measurements of silver in a transverse fieldBx5200 mT before and immediately after reversal of the lon-gitudinal field Bz56400 mT: s, absorption spectra; d , emis-sion spectra. The initial and final polarizations, calculated fromthese spectra, were 0.73 and 20.64, respectively. Note that thepeak frequency is higher when T,0 because of the strongferromagnetic susceptibility at negative temperatures. FromVuorinen (1992).

FIG. 23. Polarization pf of nuclear spins in silver, after rapidfield reversal, as a function of the initial polarization pi . Thesolid line was calculated from a theoretical model describedbriefly in the original paper. The dashed line indicates the idealcase with no loss of polarization. From Hakonen and Vuorinen(1992).

Page 31: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

31A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

in Secs. II.B, II.C, and IV.F. During the field flip, onlythe Zeeman temperature is reversed; the temperature ofthe interaction reservoir Tss is not changed. An irrevers-ible entropy increase results when HZ warms the inter-action reservoir Hss8 to negative temperatures. The cal-culation of this effect is similar to that described in Sec.IV.F. The equilibrium temperature Tf after a rapid fieldflip from B51B z to B52B z is (Slichter, 1990)

Tf52Ti~B21B loc2 !/~B22B loc

2 ! , (41)

where Ti is the initial temperature before the field flip.According to the theoretical expression of Eq. (41),

the polarization loss should be only 1.5% for the 400-mT flipping field used in the experiments. This is lowerthan the polarization losses found in actual experiments.In fact, it is not correct to use B5400 mT in Eq. (41) forsilver, since thermal mixing is exceedingly slow in thishigh field (Oja, Annila, and Takano, 1990). According tothe measured mixing times in silver, thermal contact be-tween the Zeeman and interaction reservoirs takes placein fields B,Bm'250 mT. Use of B5250 mT in Eq. (41)results in a 4% polarization loss, which is close to theobserved low-p limit.

The decrease in the flipping efficiency with polariza-tion can be understood in terms of a model that ac-counts for the fact that the high-T approximation over-estimates the heat capacity of the Zeeman system forhigh p (Hakonen and Vuorinen, 1992). The large scatterof the data in Fig. 23 suggests, however, that somethingin the experimental conditions makes the efficiency ofthe field reversal somewhat irreproducible. It is there-fore possible that saturation of the inversion efficiencyat high polarizations is not fully an intrinsic property ofsilver.

In rhodium the decrease in flipping efficiency wasmore dramatic than in silver (Hakonen, Vuorinen, andMartikainen, 1993; Vuorinen, Hakonen, Yao, and Lou-nasmaa, 1995). As a result, polarizations at T,0 werelimited to p'20.60. Field flipping and negative tem-peratures in rhodium will be discussed in greater detailin Sec. X.

Negative spin temperatures in silver have also beenproduced by applying a 180° NMR tipping pulse (Ha-konen and Yin, 1991). This technique was less successfulthan rapid field inversion. The problem with the pulsetechnique is that eddy currents shield the inner parts ofthe sample from the tipping field, making it impossibleto apply a uniform 180° pulse.

V. SUSCEPTIBILITY AND NMR DATA ON COPPER

A. Susceptibility and entropy at B50

The behavior of the static susceptibility x8(0) and en-tropy of a polycrystalline copper sample were investi-gated as a function of temperature in four series of ex-periments in Helsinki (Ehnholm, Ekstrom, Jacquinotet al., 1979; Huiku and Loponen, 1982; Huiku, Jyrkkio,and Loponen, 1983; Huiku et al., 1986). Here we de-scribe the results of measurements made between 1982

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

and 1986, using the same specimen. The sample con-sisted of eight foils, 125 mm thick and 5 mm wide; theeffective mass of the specimen was 0.035 mol, i.e., about2.5 g. The foils were welded to the first nuclear stage.After internal oxidation of magnetic impurities, the ratior5t1(15 mT)/t1(0)=2.6 was measured. The residual re-sistivity ratio increased from 260 to 8500 owing to theheat treatment; the effective RRR was about 5000. Onlythe transverse coil geometry [see Fig. 16(b)] was em-ployed in the measurements; the highest sensitivity ofthe detection system was in the y direction, perpendicu-lar to the external field and parallel to the 5-mm side ofthe foil bundle.

The measured quantity was actually the dynamic sus-ceptibility x(f)5x8(f)2ix9(f) at f510 Hz. At this lowfrequency, however, x9(f) is so small that it can be ne-glected and the measurement yields the static suscepti-bility x8(0) [see Sec. V.E for the frequency dependenceof x9(f)]. The excitation field was typically on the orderof 10 nT. An rf SQUID was used to measure the suscep-tibility signal as described in Sec. III.E.

The susceptibility vs time curve in zero field is sche-matically illustrated in Fig. 24. Following demagnetiza-tion, x8(0) first increases and, after reaching a broadmaximum, the relaxation settles roughly to an exponen-tial decrease. The overall behavior is analogous to thatobserved in electronic antiferromagnets (Kittel, 1971).The experimental discovery (Huiku and Loponen, 1982)of antiferromagnetic ordering in copper was expected,because of the negative sign of the Ruderman-Kittel in-teraction parameter R [see discussion after Eq. (5)] andtheoretical calculations based on the mean-field approxi-mation (Kjaldman and Kurkijarvi, 1979).

A characteristic metastability accompanied the transi-tion. When demagnetization from a 1-mT field wasstarted with somewhat increased entropy, the suscepti-

FIG. 24. Schematic behavior of the static susceptibility of cop-per nuclei after demagnetization to zero field. Entropy S andtemperature T are increasing with time. The maximum inx8(0) reveals a phase transition, in analogy with electronicantiferromagnets. AF1 is the ordered nuclear antiferromag-netic region, and P denotes the disordered paramagneticphase. The metastable behavior of x8(0), when demagnetiza-tion to zero field was started from a somewhat higher initialentropy, is shown by the dashed line. From Huiku et al. (1986).

Page 32: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

32 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

bility often did not follow the ‘‘curved’’ behavior buthad a higher, exponentially decaying value, characteris-tic of the paramagnetic regime. The metastability wasalso accompanied by some nonadiabaticity. These ef-fects will be discussed in Sec. V.C.

The static susceptibility of a polycrystalline coppersample as a function of temperature down to TN wasmeasured most extensively, and probably most accu-rately as well, in the work described by Huiku et al.(1986). Figure 25 shows the experimental x8(0)21 vs Tdependency in zero field. The data have been reduced tospherical sample geometry by using the equation

xsph8 ~0 !215x8~0 !2111/32Dy , (42)

where Dy50.09 is the demagnetization factor of thespecimen. The inset shows the susceptibility at high T ,together with a calculation using the high-temperatureexpansion (HTE; Niskanen and Kurkijarvi, 1981). Thedata obey the Curie-Weiss law 1/x85T/C1D down to500 nK, with C=554 nK and Q52CD= 2215 nK. TheRuderman-Kittel exchange parameter then has thevalue R52D520.39. This agrees well with R5 –0.4260.05, obtained from NMR measurements(Ekstrom et al., 1979). Below about 500 nK, the experi-mental points depart from the high-temperature Curie-Weiss behavior. The deviation can be attributed to spinfluctuations (Kumar et al., 1980). A more detailed com-parison with the various theoretical calculations is de-ferred to Sec. XV.C.4.

The static nuclear susceptibility of copper as a func-tion of entropy has been plotted in Fig. 26. The relation

FIG. 25. Inverse of susceptibility xsph8 of copper, reduced for a

spherical sample, as a function of temperature when B50.The straight line shows the Curie-Weiss behavior. The theo-retical curves refer to the high-temperature expansion (HTE,inset; Niskanen and Kurkijarvi, 1981), the spherical model(SM; Kumar et al., 1980; Lindgard et al., 1986), and the linked-cluster expansion (LCE; Niskanen and Kurkijarvi, 1983).Modified from Huiku et al. (1986).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

between x8(0) and S was found simply by a series ofdemagnetizations to B50. It was observed that a de-magnetization ends in a metastable state when the initialentropy, before the final field sweep down, is in therange between (0.3660.02)R ln4 and (0.6160.03)R ln4. Above Si50.61R ln4, the relaxation ofx8(0) is exponential and the demagnetization adiabatic,which are characteristic features of the paramagneticstate. Therefore the higher critical entropy for the first-order transition in zero field is Sc25(0.6160.03)Rln4.

The lower critical entropy Sc1 can be found by usingthe fact that the temperature is constant in the transitionregion and by assuming that the maximum of x8(0) isobtained in the antiferromagnetic state immediately be-low Tc . The assumption is supported by the observationthat electronic antiferromagnets, with an fcc lattice, usu-ally have their transition points very close to the maxi-mum of x8(0) (Domb and Miedema, 1964). Therefore,since the paramagnetic x8(0) immediately above Tc issmaller than the antiferromagnetic x8(0) at Tc , themaximum of the x8(0) vs S curve (see Fig. 26) corre-sponds to the low-entropy end of the coexistence region.This gives Sc15(0.4860.03)R ln4.

In Fig. 27 the entropy vs temperature curve is shownfor copper in the paramagnetic region below 1 mK. Thehorizontal arrow pointing to the S/R ln4 axis indicatesthe experimentally determined critical entropySc250.61R ln4 for spontaneous nuclear ordering in zerofield.

Comparisons with theory are also shown in Fig. 27.High-temperature expansions (HTE), with fifth-degree

FIG. 26. Static nuclear susceptibility of copper in zero externalfield as a function of entropy. The data have been reduced tospherical sample geometry by using Eq. (42). The triangleswere measured in a metastable state, and the circles in theantiferromagnetic phase AF1, in the paramagnetic phase P, orat T5Tc in the transition region. Sc1 and Sc2 are the lower andupper critical entropies, respectively, for the first-order phasetransition. Demagnetizations with initial entropies above Sm

ended in the metastable state (dashed lines). From Huiku et al.(1986).

Page 33: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

33A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

terms in 1/T included, can describe the experimentaldata with reasonable accuracy down to 400 nK (Nis-kanen and Kurkijarvi, 1981). The agreement improvesconsiderably when the spherical model (SM) is used(Kumar et al., 1980; Kjaldman et al., 1981). However, inview of the good agreement between the measuredx8(0) and the SM prediction, discrepancies for entropyat the lowest temperatures are puzzling. The experimen-tal S vs T data agree best with calculations using thelinked-cluster expansion (LCE) technique (Niskanenand Kurkijarvi, 1983).

In Fig. 28 the entropy curve for copper is illustratedbelow T5150 nK. The vertical line emphasizes thephase change at TN558 nK. The latent heat of the tran-sition in zero field is L5TN(Sc22Sc1)50.09 mJ/mole.The experimental points show the rapid reduction of en-tropy below TN .

B. Thermometry

Although the techniques for measuring the tempera-ture T of copper nuclei were essentially the same asthose described in Sec. IV, there were two importantdifferences. As the sample had a very high electricalconductivity, it was necessary to give the heat pulseDQ5pfB1

2x9(f)Dt/m0 using an ac magnetic field of lowfrequency to ensure full penetration. Direct measure-ment of x9(f) at f550 Hz was not feasible, however,owing to the small signal at low frequencies. Thereforex9(f) was deduced from the measured x8(0) using arelationship between x9(f=50 Hz) and x8(0) that hadbeen established from earlier NMR measurements by

FIG. 27. Nuclear-spin entropy of copper. The horizontal axisat the bottom and the vertical axis on the right show T andDs512S/R ln4, respectively, on a logarithmic scale for ex-perimental points at the lowest temperatures. The curves showthe results of theoretical calculations: HTE, high-temperatureexpansion; LCE, linked-cluster expansion; SM, sphericalmodel (see text). The linear scales at top and at left refer to theuppermost curve. Modified from Huiku et al. (1986).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Ehnholm et al. (1980). It was estimated that the errorcaused by this approach did not exceed 10%. For thedata in the ordered state, however, the error in T can belarge since the shape of the NMR signal changes(see Fig. 32) and hence also the relationship betweenx9(f550 Hz) and x8(0).

Another special problem of measuring T in copperwas associated with the data in the ordered phase. It wasobserved that demagnetization to zero field resulted inan entropy increase, to be discussed in the following sec-tion. To overcome the difficulty, T was determined inthe ordered state by using x8(0) as the thermodynamicparameter and writing T5(dQ/dx)(dS/dx) withx5x8(0). The absolute temperature could then be mea-sured in two steps. First, dQ/dx was found by applyingseveral heat pulses during the warmup at zero field. Sec-ond, S as a function of susceptibility was measured bysweeping the field at different stages of the warmup inzero field, i.e., at different values of x8(0) up to 1 mT tomonitor the polarization. The procedure was feasible asthe entropy was found to increase only during a down-ward sweep. As a result, thermometry at B50 was pos-sible although the downward sweep was nonadiabatic.

A drawback of using x8(0) as a thermodynamic pa-rameter is that the method loses sensitivity when x8(0)changes only a little upon warming. This is the case nearthe transition when x8(0) reaches its maximum value.

Thermometry was based on runs in which the demag-netization to zero field was performed for an initial en-tropy lower than 0.36R ln4 to avoid problems associatedwith metastable states (see below).

FIG. 28. Entropy of copper in zero external magnetic field as afunction of temperature below 150 nK. At the time, Tc558nK was the lowest transition temperature ever observed ormeasured. The vertical line was drawn to emphasize the first-order transition. The critical entropies are indicated by arrows.From Huiku et al. (1986).

Page 34: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

34 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

C. Metastability and nonadiabatic phenomena

In measurements on a polycrystalline copper sample(Huiku, Jyrkkio, and Loponen, 1983), a characteristicmetastability was observed after the magnetic-fieldsweep from B51 mT to zero. The behavior of x8(0)after such a sweep is illustrated in Fig. 29 for severalvalues of the initial entropy Si at B51 mT. A metasta-bility was evident when Si was larger than 0.36R ln4 butsmaller than 0.61R ln4. Then x8(0) followed the behav-ior illustrated with the dashed lines. The static suscepti-bility was always larger than x8(0) measured with initialentropies lower than Sc250.36R ln4, which was be-lieved to correspond to the equilibrium behavior.

A large irreversible entropy increase was observed, aswell, after demagnetization from Si lower than0.36R ln4. When B was swept to zero and then, after afew tens of seconds, back to 1 mT, an entropy increaseof DS50.12R ln4 was observed. The nonadiabaticitywas constant within the experimental accuracy (0.5% ofDS). The reverse sweep upward to 1 mT was adiabatic.Thus DS50.12R ln4 was added to all Si values below0.36R ln4 in order to obtain the actual spin entropy af-ter demagnetization. In experiments during which meta-stable states were reached, DS was always smaller than3%, and it vanished at 0.61R ln4.

The time scale for the irreversibility to occur was onthe order of seconds; with a stay of 2 s at B50, themeasured nonadiabaticity was only 0.06R ln4. It wassuggested that the irreversibility is intimately associatedwith the ordering process. The minimum susceptibilityafter demagnetization was reached approximately 15 safter the external field had been reduced to zero (seeFig. 29). The behavior of x8(0) and the development ofthe entropy loss both reflect a nucleation of the orderedphase from the supercooled nonequilibrium state.

FIG. 29. Static susceptibility x8(0) of nuclear spins in copperas a function of time after final demagnetization to zero field,performed from different initial entropies Si . Here S is thefinal entropy in zero field, given as percentages of R ln4.Dashed lines indicate metastable behavior. The susceptibilitydata correspond to the demagnetization factor Dy50.09. FromHuiku et al. (1986).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Subsequent susceptibility and neutron-diffractionmeasurements on a single-crystal specimen have givenmore insight into the process by which irreversibility oc-curs. The entropy increase DS50.12R ln4 was found tooccur in a field around 0.1 mT (Huiku, Jyrkkio et al.,1984; Huiku et al., 1986; Annila et al., 1992), where aphase transition also takes place from the antiferromag-netic phase AF2 to AF1; see the phase diagram of Fig. 1.Neutron-diffraction experiments (Annila et al., 1990)showed that in this region the high-field type-I ordertransforms into the (0 2

323) order. It was also directly

seen that the ordering process is associated with kineticphenomena within a time scale on the order of seconds.In the single-crystal experiments no metastable stateswere observed. Therefore it was concluded that grainboundaries in a polycrystalline sample slow down thenucleation of the ordered state, but that at entropiesbelow 0.36R ln4 a forced nucleation has to take placeowing to the absence of metastability.

D. Field effects on a polycrystalline sample

The time dependence of the static susceptibility wasalso investigated by varying the external field at whichthe final demagnetization ended (Huiku and Loponen,1982). The scheme for these measurements is schemati-cally illustrated in Fig. 30(a). According to the mean-field theory for antiferromagnetic type-I order (see Sec.XV.B.5), adiabatic demagnetization in the ordered statefollows a vertical path in the B-T plane and thereforedoes not produce any further cooling.

The susceptibility was recorded in the final fieldB5Bf during warmup. Figure 30(b) illustrates the threequantities used to analyze the data: the susceptibility im-mediately after demagnetization to the final field, x init8 ,the maximum susceptibility xmax8 , and the difference

FIG. 30. Adiabatic demagnetization into an antiferromagneticstate: (a) Schematic illustration. The demagnetization isstopped at a final field B5Bf , and the static susceptibility x8 isrecorded during warmup. (b) Definition of the three quantitiesimportant for characterizing the x8 data.

Page 35: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

35A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Dx85xmax8 2x init8 . In the beginning of the warmup, be-fore reaching the state with maximum susceptibility, thenuclei were assumed to be in the antiferromagneticstate.

The measured data are shown in Fig. 31. In low fieldsthe increase of x8(0) is largest. As a function of the finalfield after demagnetization, Dx8 first decreases until, atabout B50.08 mT, a minimum is reached. Around B50.15 mT, Dx8 has again grown to about half of thevalue observed when B50. At still higher fields the ini-tial increase decreases until, at B50.24 mT, Dx8 finallydisappears.

The anomaly at B50.08 mT indicates a magnetictransition between two ordered states, which werenamed AFl and AF2 (Huiku and Loponen, 1982). Sub-sequent experiments (Huiku, Jyrkkio et al., 1984) on asingle-crystal specimen, during which the longitudinalsusceptibility was measured as well, revealed a third an-tiferromagnetic phase, AF3. Recent neutron-diffractionexperiments (Annila et al., 1992), also performed on asingle-crystal specimen, have further shown that the an-tiferromagnetic ordering vectors, and hence also thespin structures, strongly depend on the alignment of theexternal magnetic field with respect to the crystallineaxes. It is therefore remarkable that the data on a poly-crystalline sample do not smear out the evidence forphase changes, especially at the AF1 ↔ AF2 transition.

E. NMR lineshapes of the ordered spin structures

The specimen employed in the NMR experiments(Huiku et al., 1986) was made from the same 125-mmthick Marz-grade copper foil that was used in the mea-surements of static susceptibility. The foils were first se-

FIG. 31. Initial susceptibility x init8 5x8(0) at t50, the maxi-mum susceptibility xmax8 , and the difference Dx85xmax8 2x init8as functions of the external field B in copper (Huiku andLoponen, 1982). The ordered regions AF1 and AF2 and theparamagnetic phase P are indicated. From Huiku et al. (1986).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

lectively oxidized and then rolled to their final thickness,varying from 20 to 30 mm. The final sample consistedof 36 foil pieces. For this specimen the valuesr5t1(15 mT)/t1(0)54 and t1(0)570 min were mea-sured, which are reasonably good for a sample made ofthin foils. Again, only the transverse coil system [see Fig.16(b)] was employed in the experiments. NMR spectrawere measured using the technique described in Sec.III.E. During the warmup period, typically 8 to 10curves were recorded before the sample entered thetransition region.

The absorptive part of the dynamic susceptibility isplotted as a function of frequency in Fig. 32 at severalfinal fields. The dashed curves show the lineshape justafter demagnetization, i.e., the NMR signal in the anti-ferromagnetic region. The solid curves were measured8–10 min after the end of demagnetization; they thuscorrrespond to high polarization in the paramagnetic re-gion.

It was found that the antiferromagnetic NMR line(AFL) differed at B50 in two ways from the paramag-netic line (PL): At about f51.5 kHz, the AFL wasclearly enhanced, whereas from 3 to 7 kHz it was againsmaller than the PL. The latter feature is responsible forthe fact that the static susceptibility x8(0), obtained fromx9(f) through the Kramers-Kronig relation [see Eq.(27)], increases with time (and temperature). This wastaken as the signature of antiferromagnetic ordering inthe measurements of static susceptibility. The increasedx9(f) around f51.5 kHz further supports this conclu-sion, since antiferromagnetic ordering is expected tochange the lineshape of the dynamic susceptibility. Dur-ing warmup, the AFL smoothly approaches the PL.

FIG. 32. NMR absorption line shapes of the dynamic suscep-tibility x9(f) for copper in several external magnetic fields:Solid curves, paramagnetic regions; dashed curves, antiferro-magnetic regions. From Huiku et al. (1986).

Page 36: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

36 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Qualitatively similar lineshapes were obtained at B50.03 mT.

At B50.05 mT the appearance of the AFL changes.Now the absorption in the antiferromagnetic state isconsiderably stronger between 3 and 7 kHz than in theparamagnetic region. Furthermore, the AFL resemblesthe PL even immediately after demagnetization. In thebeginning of the warmup, x9(5 kHz) is almost 1.5 timeslarger than in zero field.

An examination of the dashed line at B50.09 mTshows interesting features: x9(10 kHz) is enhanced and,at lower frequencies, x9(f) is small, which again indi-cates that the static susceptibility is rising as a functionof time. At B50.16 mT, the increase in x9(10 kHz)exceeds that of the paramagnetic line. During warmup,the AFL shifts towards the PL. The absorption at lowfrequencies is again small. At B50.24 mT, the line-shape does not change at low entropies. This shows thatthe spin system is paramagnetic immediately after de-magnetization.

It is also worth noting that the paramagnetic ‘‘back-ground’’ is large for all NMR lines measured in theantiferromagnetic phase and that the changes duringwarmup are relatively small. However, this is in agree-ment with the fact that the increase in the static suscep-tibility is small, too. A more thorough discussion of theNMR lineshapes at different fields is given in the origi-nal publication (Huiku et al., 1986). It was concludedthat the observed changes are strong evidence for mag-netic ordering in copper nuclei.

Recent theoretical calculations by Heinila and Oja(1995) have reproduced quite successfully the observedNMR lineshapes of copper in zero field (see Sec. XV.Eand Fig. 121 below).

F. Susceptibility data of a single-crystal specimen

To obtain more information about the orderednuclear-spin structures in copper, the Helsinki groupconducted experiments on a single-crystal sample byagain investigating the static susceptibility (Huiku,Jyrkkio et al., 1984; Huiku et al., 1986). In contrast to allearlier work, x8(0) was now separately measured in thethree Cartesian directions x , y , and z . In addition to theexcitation and pickup coils in the longitudinal z direc-tion [see Fig. 16(a)], there were saddle-shaped excitationcoils in both the x and y directions [see Fig. 16(b)]. Thesame saddle-shaped pickup coil was used to monitor theresponse in the latter two directions. The transverse(x ,y) and longitudinal (z) signals could be measuredduring the same run.

The specimen was a natural copper single crystal ofdimensions 0.535320 mm 3 along the x , y , and z axes,respectively. The sample material was of 99.99% purity,with the following magnetic contaminants: Mn<0.1 ppm,Cr=2 ppm, Ni<0.4 ppm, and Fe=3.0 ppm. After oxida-tion for 45 h, the residual resistance ratio (RRR) was1500. The single crystal was again connected to the firstnuclear stage by diffusion welding. For this specimen,r5t1(15 mT)/t1(0)52.8, which shows that all impuri-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

ties were oxidized. Thermal contact was found to be ex-cellent, and one could reach Te550 mK in the specimen.By x-ray-diffraction measurements it was found that thecubic (1 0 0) crystal axis was about 4° above the xyplane and 13° off of the x axis and that the (0 0 1) axiswas about 8° off of the z axis. As the external field wasaligned along the z direction, B was also aligned rela-tively accurately along the (0 0 1) crystalline axis.

In order to compare the susceptibilities in the threemutually orthogonal directions, Huiku and co-authorshad to apply a shape correction for the internal fields[see Eq. (34)] to the measured xa8 (0) (a5x ,y ,z), using

xsph,a8 ~0 !215xa8 ~0 !211~L2Di1R ! , (43)

where xsph,a8 (0) is the susceptibility related to a sphericalsample. This correction reduces the measured xa8 (0) tothe susceptibility of a virtual spin system with no mo-lecular fields arising from spin-spin interactions. In therest of this section, we use the shorter notationxsph,a8 (0)5xa . Di was estimated from the dimensions ofthe specimen, assuming an ellipsoid of similar shape tothat of the sample.

In Fig. 33 the time dependence of xa is shown for thex , y , and z directions at B50, 0.15, and 0.20 mT, respec-tively; these characteristic plots correspond to differentordered regions. The existence of the third antiferro-magnetic phase, AF3, is based on the behavior of thelongitudinal susceptibility, which now was measured forthe first time. This phase was found in fields above 0.18mT.

In zero external field, a clear increase of xa was ob-served only in the y direction; this effect is about 10%.Surprisingly, the other transverse susceptibility, xx ,

FIG. 33. The susceptibility xa of a copper single crystal alongthe three Cartesian directions (a5x ,y ,z), as a function oftime and in three external fields, B50, 0.15, and 0.20 mT.These characteristically different behaviors indicate three an-tiferromagnetically ordered regions in the nuclear-spin systemof copper. The suggested spin arrangements, based on an as-sumption of two sublattices, are also shown. From Huiku et al.(1986).

Page 37: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

37A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

shows a long plateau, whereas a small increase in xz wasseen. For B50.15 mT, the initial value of the longitu-dinal susceptibility was roughly 20% smaller thanin zero field, and xz increased at the beginning of thewarmup by almost 15%. The increase in the y directionwas not as great, but roughly the same as when B50.Again, xx stayed almost constant for the first 5 min. Inthe last plot, with B50.20 mT, a characteristic changecan be seen for xz in comparison with the behavior ob-served at B50.15 mT. Now xz decreases all the time,whereas an increase and a plateau for xy and xx , respec-tively, were found in the transverse directions.

Tentative suggestions for the spin arrangements inthese phases were made by using a simple two-sublatticemodel and an analogy with electronic antiferromagnets(Huiku, Jyrkkio et al., 1984): When x8(0) is measuredperpendicular to the sublattice magnetization, the sus-ceptibility stays roughly constant below Tc , whereasparallel to the sublattice magnetization, x8(0) ap-proaches zero as T→0 (de Klerk, 1956; Kittel, 1971). Inthe phase AF1 (B50), the staggered magnetization isthen mainly along the y axis. In the second antiferro-magnetic state, AF2 (B50.15 mT), the sublattice mag-netization has its largest component in the z directionand a smaller component in the y direction. In AF3, thespins most probably are leaning toward B5Bz. In con-trast to the ‘‘paramagnetic’’ behavior of xz , a small in-crease in xy indicates a staggered magnetization parallelto y in this phase. These suggestions for the antiferro-magnetic structures are schematically illustrated in Fig.33.

More of the single-crystal data are illustrated inFig. 34, where the net rise of susceptibility, Dxa

5xamax2xa

init , has been plotted as a function of the ex-ternal field at which the final demagnetization wasstopped (Huiku et al., 1986). In low magnetic fields,Dxy is roughly 10% of xy

max ; the value of Dxy drops toabout 2% for fields between 0.06 and 0.10 mT. FromB50.13 to 0.16 mT, Dxy is almost as large as in zerofield. After B50.16 mT, a small sudden drop can beseen. With increasing field, Dxy decreases to zero atabout B50.25 mT.

Data in the x direction are in sharp contrast: Overmost of the region from B50 to 0.25 mT, Dxx50. It ishard to know whether the increase around B50.14 mT,supported essentially by one data point in Fig. 34, is realor not. The large susceptibility anisotropy between thex and y directions is puzzling. It was suggested (Huikuet al., 1986) that the effect is caused by the slab shape ofthe specimen, which is sensed by the anisotropic dipolarinteraction.

The data in the longitudinal geometry, illustrated inthe lower part of Fig. 34, show that Dxz is small below0.08 mT. In higher fields Dxz increases rapidly and isroughly 12% of xz

max between B50.11 and 0.16 mT. Instill higher fields, Dxz decreases quickly and vanishes atB50.20 mT, in contrast to Dxy , which disappears onlyat B50.25 mT.

From the results illustrated in Fig. 34 and from a care-ful analysis of the original data (Huiku et al., 1986), the

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

following picture emerges (see Fig. 1): The first antifer-romagnetic phase, AF1, is stable below B50.06 mT.Between 0.06 and 0.11 mT there is a transition region.The second phase, AF2, is stable from B50.11 mT toB50.16 mT. The transition from AF2 to AF3, between0.16 and 0.18 mT, is marked by a large change in Dxzand by a small jump in Dxy . The AF3 phase existsabove B50.18 mT and below B50.25 mT. The transi-tion to the paramagnetic state proceeds by tilting thespins more and more toward B until, at the critical fieldBc50.25 mT, AF3 and the paramagnetic phase P arethe same.

The interpretation of the susceptibility data and theexistence of the AF2 phase between AF1 and AF3 showthat all the ordered spin structures have different sym-metries, and therefore we are led to suggest that thetransitions between them are of first order. BetweenAF1 and AF2 this conclusion is supported by the largenonadiabaticity, DS50.12R ln4, and by the metastabili-ties, discussed in the previous section. The first-ordernature of the transition between AF2 and AF3 is basedon the rapid change in Dxz and on an analogy with spin-

FIG. 34. Normalized net rise in the susceptibility of copper,Dxa /xa

max , plotted as a function of the external field B5Bz;xa

max is the maximum susceptibility. Data for the transversedirections are plotted in the upper frame @a5x (d); a5y(s)], and for the longitudinal z direction below. From Huikuet al. (1986).

Page 38: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

38 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

flop transitions in electronic systems (De Klerk, 1956).The transition from AF3 to P is of second order.

G. Magnetization

Magnetization has not been measured directly in thecourse of the Helsinki studies on copper. In principle,the magnetization M can be calculated by integratingthe measured longitudinal susceptibility over the fieldB . Such a procedure was first performed by Annila et al.(1992) to analyze their neutron-diffraction data. It wasnecessary to know M to be able to correct for the de-magnetization field, which is given by the shape-dependent demagnetization tensor D and M, viz.,

B5Be1Bd5Be2m0DM. (44)

Here B is the magnetic field in the sample, and B e is theexternal field. In the earlier susceptibility measurementson copper, B e was always applied along the long axis ofthe sample, i.e., the z axis. In this case, B d vanishesbecause Miz and Dzz'0. Therefore B5Be . In the workof Annila et al. (1992), however, Be was in some mea-surements applied almost along the hard direction ofmagnetization, which corresponds to a large demagne-tizing factor (see Fig. 53 below for the geometry), so thatBd was even larger than Be .

Magnetization for a certain field and entropy can becalculated by remembering that the measured xL is theadiabatic susceptibility m0(dM/dB)S and by assumingthat the demagnetization to zero field takes place at con-stant entropy. To obtain, for example, M(B) immedi-ately after demagnetization, one can integrate xL mea-sured at t50,

M~B !5E0

BcxL~B ,t50 !dB/m0 . (45)

The demagnetization to B50 is not, however, adiabaticbecause of the entropy increase in fields around 0.1 mT.This makes the integration somewhat inaccurate.

Annila et al. (1992), nevertheless, used Eq. (45) to ob-tain an approximate M . A particularly interesting quan-tity is the magnetization at the critical field immediatelyafter demagnetization. The xL data, calibrated at B50using the results of Huiku et al. (1986), yieldM50.57Msat at B5Bc50.25 mT. This value may seemsmall in view of the fact that the mean-field theory pre-dicts M50.9Msat for the estimated entropy after demag-netization. The discrepancy can be explained by quan-tum effects and/or substantial antiferromagnetic short-range correlations.

The actual model used by Annila et al. (1992) to cor-rect for B d was, in fact, simpler than that obtained fromEq. (45). In the whole ordered region it was assumedthat

M5xB/m0 , (46)

with x53.1. This approximation seems reasonable be-cause the variation in xL(B), measured immediately af-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

ter demagnetization, is only 14%. Furthermore, themean-field theory for a type-I spin arrangement alsopredicts a temperature- and field-independent suscepti-bility in the ordered region (Kumar et al., 1986). Equa-tion (46) with x53.1 gives M50.61Msat at Bc in goodagreement with the above value obtained by integration.Further support for this model is given by neutron-diffraction measurements in which the intensity of the(1 0 0) Bragg reflection was measured in two symmetri-cally equivalent directions, namely, for B along the [011]and [01 1] axes. It was observed that the behavior of the(1 0 0) intensity as a function of B was similar in bothdirections, in spite of the different demagnetization cor-rections D=0.83 and 0, respectively [see Figs. 57(a) and(b)]. It was concluded that Eq. (46) gives M correctlywithin 15%.

H. Entropy diagram

By studying the susceptibilities xy and xz as functionsof the entropy Si just prior to the final demagnetization,it was possible to determine the magnetic field vs en-tropy diagram of the nuclear-spin system in copper(Huiku et al., 1986). Si was controlled by varying thedemagnetization procedure: One simply had to wait fordifferent lengths of time at B51 mT before sweepingthe magnetic field down to its final value; Si could becalculated from t1 and the waiting time.

The B-S phase diagram for the nuclear-spin system incopper was already illustrated in Fig. 1; it was con-structed by demagnetizing from different initial valuesof entropy, thus moving vertically down, and then byallowing the specimen to warm up, thus advancing hori-zontally to the right while the susceptibility was beingmeasured. The shadowed areas indicate regions whereone characteristic behavior of susceptibility changes toanother because a first-order phase change is proceedingin this area. During transitions, neighboring phases co-exist as macroscopic domains. The direction of the ex-ternal field deviated 8° from the [001] crystalline axis.The critical field Bc50.27 mT can be found by extrapo-lating the AF3 – P boundary to S50; the value of Bc isslightly higher than that deduced in Sec. V.D.

The relation between entropy and temperature wasnot determined for the single-crystal specimen. Accord-ing to data on the polycrystalline sample, however, thetransition region in zero field corresponds toTc558610 nK (see Fig. 27).

VI. EXPERIMENTAL TECHNIQUESOF NEUTRON-DIFFRACTION MEASUREMENTS

A. Setup for neutron-diffraction experiments

The neutron-diffraction experiments on copper (Jyrk-kio et al., 1989; Annila et al., 1992) were carried out inthe guide hall adjacent to the DR-3 reactor at the RisøNational Laboratory in Denmark. From a cold source, abeam of long-wavelength neutrons was transmitted via acurved guide to the TAS-8 double-axis diffractometer,

Page 39: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

39A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

constructed at the Hahn-Meitner Institute in Berlin.This arrangement reduced mechanical vibrations as wellas electrical noise, and it provided a clean beam, freefrom g rays and fast neutrons. The background radia-tion in the experimental hall was also very low. Furtherattenuation was obtained by decoupling the cryostatfrom the building and the pumping systems by effectivevibration isolators and by employing extra electricalshields and filters for all current lines entering the refrig-erator (see Sec. VI.B).

The diffractometer could be used in a polarized or anunpolarized mode. The setup for polarized neutrons(Jyrkkio, Huiku, Clausen et al., 1988) will be discussedin Sec. VI.D. Figure 35 shows schematically the experi-mental arrangement for unpolarized neutrons.

A graphite monochromator was used to select the de-sired wavelength from the neutron beam. Usually l= 4.7Å was chosen, but sometimes l=2.4 Å was used. Theflux at 2.4 Å was about 83105 neutrons/cm 2 s, deducedfrom activation of a gold foil at the sample position.

In addition to the fundamental wavelength, the beamreflected from the monochromator contains neutronswith wavelengths l/2, l/3, etc. To obtain a fully mono-chromatic beam in an actual experiment, the higher-order contaminations were filtered out. A pyrolitegraphite filter at l=2.4 Å and a liquid-nitrogen-cooledBeO filter at l=4.7 Å were employed. The higher-orderreflections were useful, however, in obtaining the correctorientation of the sample and the detector before anactual experiment. For example, the lattice reflection(2 0 0) for the wavelength l/2 is found at the same posi-tion as the antiferromagnetic Bragg reflection (1 0 0) forthe wavelength l . For this reason, an effective filteringof the l/2 neutrons is needed during the ordering ex-periment.

In the first measurements, a single detector was used.During later experiments a horizontal, position-sensitivelinear detector was employed as well. This was used to

FIG. 35. Schematic top view of the neutron-diffraction setupat Risø. The cryostat can be rotated on the spectrometer turn-table, and the detector can be positioned at the desired scat-tering angle. From Jyrkkio et al. (1989).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

study the profile of the Bragg peak, since mechanicalmovements of the spectrometer during a lineshape scanmight have heated the sample. Later it was found, how-ever, that conventional scanning of reciprocal space canbe used even when the nuclei are in the ordered statebelow TN558 nK (Annila et al., 1990). Magnetic fieldson the sample and on the first nuclear stage are then lowand vibrational heating due to eddy currents is small.Scanning shortened the time spent in the ordered stateby about 40% compared with experiments without me-chanical movements. This reduction in the availablemeasurement time was large but acceptable.

Both the incoming and scattered beams were effec-tively masked, using boron-containing plastic plates tominimize the background scattering seen by the detec-tor. With a scattering angle close to 90°, the backgroundcount for the single detector was about 0.3 neutrons/sand for the linear detector about 0.7 neutrons/s.

B. Risø cryostat

The design of the Risø cryostat (Jyrkkio et al., 1989) isillustrated in Fig. 36. The apparatus consists again of ahigh-power dilution refrigerator and two copper nuclearstages, working in series. The operating principles arebasically the same as for the Helsinki apparatus, de-scribed in Sec. III.B. To stop vibrations from enteringthe cryogenic system, the spectrometer was decoupledfrom the floor by spring-type vibration isolators.

The Risø cryostat was standing on a turnable spec-trometer so that the sample could be oriented as de-sired. The surroundings of the specimen were designedto give maximum openings for the incoming and scat-tered neutron beams, with only minor attenuation.Therefore the solenoid used for demagnetizing thesample was a split-pair superconducting coil, producing4.6 T at its center. The two halves of the magnet wereseparated by three aluminum wedges that covered about45° of the scattering plane. The field of the split-pairmagnet was asymmetric in such a way that there were nofield-free volumes in the beam region. This is importantfor experiments with polarized neutrons, since a zero-field region would cause depolarization of the neutronbeam.

The commercial Oxford 600 dilution unit (Oxford In-struments, Eynsham, OX8 1TL, England) was equippedwith a continuously filling condenser. All pumping linesconnected to the top of the cryostat were flexible toavoid transmission of vibrations to the apparatus. Thefirst nuclear cooling stage was again made of a bulkpiece of copper with vertical slits. Its weight was 2.5 kgbut the effective mass in the 8-T field of the demagneti-zation magnet was only 1.0 kg, since large sections of thenuclear stage were located in the field-compensated re-gions. The heat switch, somewhat different from that inthe Helsinki machine, was constructed of zinc foils hav-ing pressed contacts to copper ‘‘fingers.’’

Page 40: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

40 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

FIG. 36. Low-temperatureparts of the two-stage nuclear-demagnetization cryostat atRisø. The whole assembly isimmersed in a liquid-heliumdewar. From Jyrkkio et al.(1989).

The immediate surroundings of the sample are shownin more detail in Fig. 37. The specimen was connected tothe first nuclear stage via a high-conductivity thermallink made of copper. There were two radiation shieldsaround the sample: the inner one was anchored to themixing chamber and the outer one to the still. A smallsymmetric split-coil magnet, wound on top of the outerradiation shield, produced the field on the sample afterdemagnetization of the main split-pair magnet. The ex-citation and pickup coils for measuring the susceptibilitywere placed below the split region of the magnet, sinceotherwise neutrons scattered from the coil system wouldhave increased the background count during experi-ments.

Cooling of the sample was performed in the same wayas in the earlier Helsinki experiments (see Sec. III.B).Small modifications were necessary, however, becauseof the larger heat capacity of the copper parts in thethermal link between the sample and the first nuclear

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

stage. This was due to the field profile of the main split-pair magnet, with the field extending higher up than inthe Helsinki apparatus.

The nuclear stages, the upper one typically magne-tized to 7.6 T and the lower to 4.4 T, were first precooledby the dilution refrigerator to 10 mK in 36 h. This lowstarting temperature was necessary for obtaining almostperfect adiabaticity during nuclear cooling. The firststage was then demagnetized to 0.1 T in 8 h. During thefield sweep, dB/dt was decreased several times to elimi-nate eddy-current heating towards the end of demagne-tization. A conduction-electron temperature of about150 mK was reached.

Next, the large split-pair sample magnet was demag-netized from 4.4 T to 2.0 T in 40 min. Immediately fol-lowing, the fields in the upper and lower magnets werereduced, in 25 min, to 50 mT and to 1.0 T, respectively.The next step was a fast demagnetization of the split-pair magnet from 1 T to zero. At the same time, the first

Page 41: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

41A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

stage was further demagnetized to 30 mT.The specimen was now in the 10-mT field produced by

the small split-pair coil inside the mu-metal shield (seeFig. 37). One then had to wait 30 s to let the remanentfield of the large split-pair magnet stabilize. The finalstep in the cooling procedure was demagnetization ofthe specimen from 10 mT to B,0.3 mT in about 20 s.The sample nuclei were thereby cooled below the anti-ferromagnetic ordering temperature TN558 nK,whereas the conduction electrons stayed at 50–100 mK.

Three different techniques were employed to deter-mine the temperature during the various stages of theRisø experiments (Jyrkkio et al., 1989). A cobaltnuclear-orientation thermometer (Berglund et al., 1972)was mounted on the first nuclear stage (see Fig. 36) tomeasure the precooling temperature. Since this ther-mometer could be used from 2 to 40 mK, it also indi-cated the temperature at the beginning of demagnetiza-tion. The cobalt thermometer was also employedindirectly for measuring the final nuclear temperature ofthe first stage, which was very close to the final Te of thesample when the specimen was not heated by the neu-tron beam. For this purpose the upper nuclear stage wasslowly magnetized from Bf530 mT to a field between 1and 2 T. Assuming an adiabatic process, the tempera-ture in the low field was simply found from the readingof the cobalt thermometer and from the ratio of thefields [see Eq. (18)]. The lowest temperatures measuredby this technique were about 70 mK.

The final conduction-electron and lattice temperaturecould also be found more directly by measuring t1 fromthe susceptibility signal and by using the Korringa rela-tion, Eq. (9). The longest relaxation times observed cor-responded to Te550 mK.

A polarized neutron-scattering technique was also de-veloped for thermometry (Jyrkkio, Huiku, Clausenet al., 1988). The important advantage of this method isthat neutrons probe the nuclei directly; the method thus

FIG. 37. Surroundings of the sample in the Risø cryostat, in-dicating the locations of the susceptibility coils, the small split-pair sample coil, and the mu-metal shield. From Jyrkkio et al.(1989).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

avoids problems associated with insufficient thermalcontact. The sensitive temperature range of this ther-mometer can be chosen at will by applying a suitableexternal magnetic field. The method was used to mea-sure the nuclear-spin temperature from 20 mK down to100 mK. We defer a thorough discussion of this versatiletechnique to Sec. VI.D.

Susceptibility measurements also played an importantrole in the Risø studies of nuclear ordering. A SQUIDwas used in the flux-locked mode (Lounasmaa, 1974),and a flux transformer was built to match the inductanceof the pickup loop to the signal coil of the superconduct-ing sensor. The positions of the SQUID, the flux trans-former, and the astatic pickup coil are shown in Fig. 36.

Figure 38 is a photograph of the experimental ar-rangement in Risø.

C. Beam heating

An early series of measurements was made at Risø todetermine beam heating of the copper sample (Jyrkkioet al., 1989). The reason for these experiments was two-fold. First, for planning the work on copper, one had toknow the expected amount of beam heating to makesure that neutrons would not prevent nuclear orderingaltogether or destroy the ordered state too fast. Second,information about beam heating was important forall future neutron-diffraction experiments on nuclearmagnetism, so that the reliability of calculated estimates

FIG. 38. Experimental setup employed in Risø for neutron-diffraction experiments on copper nuclei. The cryostat ismounted on the spectrometer turntable. The single-channelneutron counter is at left. Risø photograph.

Page 42: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

42 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

could be assessed. Beam heating manifests itself by rais-ing the electronic temperature of the sample. This, inturn, speeds up the spin-lattice relaxation process [seeEq. (9)] and thereby warms the system faster across theorder-disorder phase boundary.

The beam heats the target mainly by processes follow-ing thermal neutron absorption but also through pos-sible contaminations of the beam by g rays and/or byepithermal and fast neutrons (Steiner, 1985). In the Risøsetup, all these effects, except the first, could be ne-glected because the curved neutron guide ensured asmall background and reduced all contaminations to avery low level.

Neutron absorption is followed by prompt g rays and,later, by b emission. The radioactive nucleus formed byneutron capture is in an excited state and decays withina few picoseconds to its ground level by emitting one ormore prompt g’s. The energy of the 66Cu excited stateabove the ground level is 7.1 MeV. Over 95% of it es-capes because the mean paths of g rays before theyleave the sample are much shorter than the absorptionlength. The energy input from this source is about 0.1nW, which should not cause serious difficulties.

Since the half-life of 64Cu is 12.8 h, the heating causedby the b decay of this isotope has no practical signifi-cance in the experiments, which typically are finished in15 min. However, the half-life of 66Cu is 5.1 min, whichmeans that its b decay must be assessed carefully. bheating is more complicated than that due to the g’ssince the b-energy distribution is continuous and be-cause the stopping-power formula in metals is relativelycomplicated (Knoll, 1968). It turns out that a large por-tion of the b energy is absorbed in the specimen andthat the shape of the sample significantly affects the out-come. By making one dimension of the specimen clearlysmaller than the others, one enables more of the b en-ergy to escape. This was the reason why slablike, 0.5-mm-thick crystals were employed in all neutron-diffraction experiments on copper. Even with thisconstruction, about 40% of the b energy was absorbedby the sample. Calculated values for the b heating are0.1 nW for natural copper and 0.6 nW for 65Cu. As ex-pected, increasing the isotopic proportion of 65Cu in-creased the b absorption significantly.

However, for reasons to be explained in Sec. VII.A,an isotopically enriched 65Cu single crystal was em-ployed in all neutron-diffraction measurements. Beamheating was accurately measured for the 65Cu sample bystudying its effect on the spin-lattice relaxation rate.This method is straightforward because it is based on theKorringa relation, Eq. (9); t1 was deduced from thex8(0) signal, measured by a SQUID. It was only neces-sary to find the change in t1 due to the neutron beam.Since the beam heating increases Te , t1 becomesshorter. The size of this effect is dependent on Qbeamand on the thermal resistance between the sample andthe first nuclear stage. The resulting temperature differ-ence is given by

Te22Te1

2 52RQbeam , (47)

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

where Te1 is the electronic temperature in the firstnuclear stage and R is the heat resistance of the thermallink.

When the beam was off, the heat leak to the samplewas very small and could be neglected. Therefore oneonly had to know R to apply Eq. (47). Experiments inzero or low field, without the beam, gave t1540 min;with a flux of 23105 neutrons/cm 2 s the relaxation timewas shortened to t1520 min. Using the measured val-ues, k50.2 sK in zero field and R511 K2/W, in Eqs. (9)and (47), Jyrkkio et al. (1989) obtained Qbeam51.0 nW.The agreement between the calculated value for b-rayheating and the experimental result is good.

D. Use of polarized neutrons

Measurements of spin polarization are an importantpart of nuclear thermometry. A conventional way tomake these measurements is first to record the NMRsignal of the sample nuclei as described in Sec. IV.A andthen to calculate p . Another technique that has recentlybeen employed to measure p makes use of polarizedneutrons (Steiner et al., 1981; Benoit et al., 1982b;Steiner, 1990, 1993); the method is practical in materialsfor which the spin-dependent part of the scatteringlength b is comparable to, or larger than, the spin-independent part b0. This, in fact, is the case for severalnuclei (Koester and Rauch, 1981). Therefore the scatter-ing cross section for thermal neutrons is strongly af-fected by the degree of nuclear polarization and by therelative orientation of the neutron and the nuclear spin.A measurement of Bragg intensities thus gives directinformation on nuclear polarization of the sample.

Measurement of p with polarized neutrons has someadvantages over the NMR method. The resonance con-dition often limits the NMR technique to certain mag-netic fields, whereas this restriction does not apply toneutron thermometry. Difficulties due to skin depth andeddy-current heating in bulk metallic samples are alsoabsent when using neutrons. Instead, one has to worryabout the energy absorbed from the beam. This heatgoes, however, primarily to the conduction-electron sys-tem, unlike the heat from an NMR tipping pulse appliedto the sample nuclei. Beam heating affects the wholesample more uniformly than heating caused by a high-field NMR measurement; in the latter case the heat isabsorbed in the skin layer. However, recording theNMR response, when feasible, is much faster than apolarized-neutron measurement.

The scattered intensities from Bragg reflections areproportional to the squares of the structure factorsuF1(p)u2 and uF2(p)u2, which are different for interact-ing systems consisting of a nucleus and a neutron in totalspin states I1 1

2 and I2 12, respectively. The structure fac-

tors are

uF6~p !u2516~b026b0bIPp1 1

4 I2b2p2!, (48)

Page 43: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

43A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

where p is the nuclear polarization and P that of theneutron beam. The ratio between the scattered intensi-ties for the two spin states is called the flipping ratio(FR),

FR5uF1~p !u2/uF2~p !u2 . (49)

This quantity is less sensitive to experimental errorsthan the absolute intensities. Furthermore, once the po-larization of the beam is known, measurement of theflipping ratio directly gives p without any further cali-brations. However, this is true only if the neutron beamis not significantly attenuated by scattering when itpasses through the sample. In real crystals correctionsdue to extinction must be taken into account. The scat-tered intensity then increases more slowly thanuF6(p)u2 and causes a decrease in the observed flippingratios relative to those calculated by Eq. (49). The mag-nitude of extinction depends on the wavelength and onthe quality of the crystal.

1. Experimental setup

The setup used at Risø to study scattering of polarizedneutrons from copper nuclei is illustrated in Fig. 39(Jyrkkio, Huiku, Clausen et al., 1988). A similar experi-mental arrangement was previously employed in theHahn-Meitner Institute to investigate nuclear magne-tism in HoVO 4 at much higher temperatures (Steiner,1985; Steiner et al., 1986). The use of polarized neutronsmakes the whole setup somewhat more complicatedthan that needed for an unpolarized beam (see Fig. 35).For a general discussion of neutron instrumentation, werefer the reader to Windsor (1986). The neutrons werepolarized after passing the graphite monochromator bya supermirror system. The polarization was approxi-mately 95% and the transmission about 10%. Thus the

FIG. 39. Neutron-diffraction setup at Risø for studies on cop-per using a polarized beam. From Jyrkkio, Huiku, Clausenet al. (1988).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

flux was considerably smaller for polarized than for un-polarized neutrons. The polarizer and the neutron guidereduced the second-order contamination of the beam toless than 1% at the wavelength l52.4 Å. Permanentmagnets were used to produce the vertical guide fieldsbefore and after the cryostat.

The neutron-beam polarization could be reversed by adc flipper, located in the guide field before the cryostat.The efficiency of the flipper was better than 99%. Thebeam polarization was monitored by an analyzer in thebeam behind the cryostat. The analyzer measured theflipping ratio of the (2 0 0) reflection from a magneti-cally saturated Co 92Fe 08 crystal.

Measurements were made by first counting with theneutron polarization up (1) and then repeating thecount with the polarization down (2); the change wasproduced by operating the flipper. The ratio of the peakcounts from the (2 0 0) reflection of copper, correctedfor the background, was then calculated. The simulta-neously measured flipping ratio at the analyzer was em-ployed to monitor the beam polarization.

2. Flipping ratio versus polarization in copper

The observed flipping ratio for a copper single crystalof natural isotopic mixture is plotted in Fig. 40 vs thenuclear-spin polarization (Jyrkkio, Huiku, Clausen et al.,1988). The measured flipping ratio is approximately30% smaller than the theoretical prediction presentedby the solid curve. The discrepancy was attributed toextinction and could be quantitatively explained interms of a theoretical correction (Zachariassen, 1967).Another noteworthy feature of the data is the interceptat p50 where FR50.98 instead of 1. This effect is dueto the different transmission coefficients for the two

FIG. 40. Observed flipping ratio vs nuclear-spin polarizationmeasured for a natural copper sample. Open symbols l52.4Å; filled symbols l53.0 Å. The solid line is for an extinction-free sample. From Jyrkkio, Huiku, Clausen et al. (1988).

Page 44: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

44 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

neutron polarizations through the saturated mu-metalshield around the sample (see Fig. 37).

The measurement was later repeated for the 65Cuspecimen as well. Equation (49) then predicts that theslope of the FR vs p curve should be 2.5 times higherthan that for natural copper. Surprisingly, however, itwas found that flipping ratio is less sensitive to nuclear-spin polarization in the 65Cu sample (Jyrkkio et al.,1989). The effect could be explained by larger extinctionin the isotopically pure specimen.

Polarized-neutron measurements were used for sev-eral purposes in the course of the studies on copper.Since the experiments were always made in a high mag-netic field, the nuclear-spin temperature could be ob-tained from the measured polarization using Eq. (25b).Spin-lattice relaxation was investigated under conditionsin which the Zeeman splitting in temperature units iscomparable to the electronic temperature of the sample.Theoretical calculations (Jauho and Pirila, 1970; Baconet al., 1972; Shibata and Hamano, 1982) then predict de-viations from the Korringa behavior t1Te5k with in-creasing B/Te . The measurements (Jyrkkio, Huiku,Clausen et al., 1988) supported the calculations, but thedata did not extend to B/Te values needed for a pro-nounced deviation from Korringa’s equation.

Polarized neutrons were also employed to determinethe thermal conductivity of the silver link between thesample and the first nuclear stage, as well as the heatleak to the copper specimen. It was also possible tomonitor directly the sample polarization during demag-netization of the first nuclear stage. This provided a use-ful check for nuclear ordering experiments. While allthese studies, in principle, could have been performedby measuring the NMR signal, assuming that difficultiesassociated with skin-depth effects would have been tol-erable, polarized neutrons provided a versatile and use-ful tool that was not sensitive to the external magneticfield.

The insensitivity to B must be taken with some cau-tion, however. It was observed (Jyrkkio, Huiku, Clausenet al., 1988) that changes in the stray fields of the twolarge superconducting magnets slightly affected the ab-solute intensities, although the flipping ratio (FR) wasalmost unchanged. More severe were the low-field re-strictions. In the work on copper (Jyrkkio et al., 1989),the field on the sample had to be at least 0.2 T. Belowthis value the mu-metal shield (see Fig. 37) was not satu-rated, and therefore it would have strongly depolarizedthe neutron beam passing through the shield. Polarized-neutron measurements in lower fields would have beenpossible if, e.g., the mu-metal had been replaced by asuperconducting shield made of aluminum.

It is not clear, however, whether measurements withpolarized neutrons are feasible down to the fieldsneeded for antiferromagnetic ordering in copper and sil-ver, i.e., in fields on the order of 0.1 mT. A small guidefield, typically on the order of 10 mT, or a strictly zerofield is needed to preserve the polarization of the neu-tron beam. In any case, the order-of-magnitude lowerneutron flux as compared with the unpolarized beam,

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

and the ensuing long counting times, make it difficult touse polarized neutrons for investigations of the orderingitself, at least in copper and silver.

VII. NEUTRON-DIFFRACTION EXPERIMENTSON COPPER

A. Neutron scattering from ordered copper nuclei

In the absence of an electronic magnetic moment, thesquared structure factor F2, to which the scattered-neutron intensities are proportional, can be written(Schermer and Blume, 1968; Moon et al., 1969; Price andSkold, 1986) as

uF~k!u25 (d ,d8

@b021 1

2 b0bP•~^Id&1^Id8&!1 14 b2^Id&•^Id8&#

3exp@2ik•~d2d8!# , (50)

where b0 is the coefficient of the spin-independent partand b that of the spin-dependent part of the neutron-nucleus scattering amplitude b01bI•s; here I is the spinof the nucleus, s is the neutron spin, k is the scatteringvector, i.e., the difference between the initial and finalwave vectors of the neutron, and P is the polarization ofthe neutron beam. ^Id& denotes the thermal average ofthe spin at site d. The sum is taken over the magneticunit cell, its size depending on the ordered structure.

The first term b02 in Eq. (50) is the normal structure

factor of an fcc lattice, being nonzero and thus produc-ing neutron peaks only for reciprocal-lattice vectors withall indices (hkl) even or all odd [(1 1 1), (2 0 0), (2 2 0),etc.]. The second term is the only one that depends onthe beam polarization P ; it plays an important role forexperiments in the paramagnetic regime (see Sec. VI.D)and it is also crucial in studies of ferromagnetic ordering.However, for antiferromagnets in zero field, (d^Id&50and the second term vanishes.

The third term in Eq. (50) is responsible for the ap-pearance of additional Bragg peaks in the antiferromag-netically ordered state (Steiner, 1993). We call this termFAF and rewrite it employing ^Id&5pdInd where pd isthe nuclear polarization8 or, in the ordered state, theantiferromagnetic sublattice polarization at site d , andnd is a unit vector in the direction of the spin at the samesite. By further assuming that pd5p for all d , one finds

uFAF~k!u25 14 b2I2p2U(

dndexp~ ik•d!U2

. (51)

This term is zero for all fcc reflections, but nonzero foradditional antiferromagnetic peaks, depending on sub-lattice spin directions nd . If the translational period ofthe antiferromagnetic unit cell is the same as the lengthof the fcc unit cell, reflections with (h k l) mixed appear

8In conventional notation used in the literature on neutronscattering, the symbols P and p are exchanged. Here, however,we follow the notation that is consistent with the rest of ourpaper.

Page 45: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

45A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

[(1 0 0), (1 1 0), (2 1 0), etc.]. For a larger magnetic unitcell or for an incommensurate structure, nonintegral in-dices become possible as well.

To offer an idea of the strength of the expected anti-ferromagnetic peaks, Table III lists constants b0, b0

2, b ,and 1

4b2I2 for natural copper and for 65Cu (Koester and

Rauch, 1981). The essential facts are that the intensity ofall antiferromagnetic peaks are smaller than the fcc re-flections by a factor of 100 or more, even in the case offull sublattice polarization, and that the use of 65Culeads to a sixfold gain in the scattered intensity com-pared to natural copper, since the spin-dependent scat-tering length of 65Cu is larger than that of 63Cu. Thisclearly favors the use of 65Cu isotope for neutron-diffraction studies.

B. Observation of nuclear magnetic ordering

Mean-field calculations (Kjaldman and Kurkijarvi,1979) for the ordered state in copper at B50 predict anantiferromagnetic type-I structure in which the magneticunit cell equals the primitive cell. The neutron-diffraction measurements were, therefore, started with asearch for the antiferromagnetic Bragg peak at the(1 0 0) position, which is the fundamental reflection of atype-I structure. The first results of these experimentswere reported by Jyrkkio, Huiku, Lounasmaa, Siemens-meyer, Kakurai, Steiner, Clausen, and Kjems (1988).For a more detailed description, we refer the reader toJyrkkio et al. (1989).

In all experiments on nuclear ordering, a single-crystal65Cu specimen was used. It was prepared by startingfrom isotopically enriched 65Cu powder, which con-tained several hundred ppm’s of the magnetic contami-nants Fe, Cr, Ni, and Mn. The material was first electro-lytically purified, which reduced the impurity level bymore than an order of magnitude. A single crystal wasthen grown in a graphite crucible, using a seed crystal toobtain the desired orientation. After the crystal wasmade, it was selectively oxidized; during this procedurethe residual resistivity ratio increased from 50 to 350.For spin-lattice relaxation, the ratio r5t1(10 mT)/t1(0)'6 was measured, showing that impurities still sig-nificantly enhanced the relaxation.

The approximate dimensions of the sample were353730.6 mm3. The [01 1] axis pointed approximatelyalong the longest edge of the specimen. After mounting,the [01 1] direction was along the applied external fieldwithin 4°, while the [100] and [011] axes were in the

TABLE III. Neutron-scattering constants for natural copperand for 65Cu.

Quantity Natural Cu 65Cu

b0 (10 214 m) 0.77 1.00b0

2 (b) 0.593 1.00b (10 214 m) 0.073 0.18514b2I2 (b) 0.0030 0.0193

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

scattering plane, i.e., in the plane perpendicular to theaxis of the sample magnet. The orientation of the speci-men was made using the (2 0 0) fcc reflection and thel/2 component of the neutron beam with l54.7 Å.

The single counter used in the early stages of thesemeasurements was soon replaced by a linear, position-sensitive detector. Special precautions were needed toobtain fields below Bc because of insufficient shieldingprovided by the mu-metal tube around the sample. Con-sequently, during the short halt of demagnetization atthe 10-mT field, generated by the small coil (see Fig. 37)inside the shield, a negative current was fed into thesplit-pair magnet to compensate for its remanent field.In later experiments, an additional mu-metal tube wasemployed, which improved the shielding considerably,and compensation was no longer necessary.

To minimize the heating caused by neutrons, thebeam port was opened only during demagnetizationfrom B510 mT to the final experimental field B,Bc ;this sweep took typically 20 s. After having reached thefinal field, the scattered intensity and the static suscepti-bility were monitored as functions of time. During the(1 0 0) experiments the magnetic field at the position ofthe susceptibility coil was about 10% less than in theupper part of the sample that was probed by neutrons.For later experiments the coil was remade so that thefield homogeneity was 65% over the entire specimen.

The most complete set of data (Jyrkkio, Huiku,Lounasmaa et al., 1988; Jyrkkio et al., 1989) at B50 isshown in Fig. 41. A clear Bragg reflection at the (1 0 0)position was observed after demagnetization to zerofield. The neutron intensity increased slightly during thefirst minute. After this the count rate decreased, indicat-ing a decay in the antiferromagnetic sublattice polariza-tion [see Eq. (31)] as the nuclei warmed up because ofthe spin-lattice relaxation process. After 5 min the de-crease became slower, and after 7–8 min no neutronswere observable above the background.

The inset in Fig. 41 illustrates the simultaneously mea-sured static susceptibility: x8(0) shows almost a plateaufor the first 4–5 min and thus also indicates antiferro-magnetic order. Later, x8(0) bends towards the para-magnetic exponential relaxation, which is obvious 7–8min after the final field was reached. The disappearanceof the neutron signal thus coincides approximately withthe onset of exponential relaxation of x8(0). Thereforethe neutron-diffraction data confirm the previous con-clusions obtained from measurements of the static sus-ceptibility (see Sec. V.A) (Huiku et al., 1986).

To prove that the scattered intensity is, indeed, aBragg peak, a position-sensitive detector was employed.The time evolution of the peak is shown in the lowerpart of Fig. 41, at four 75-s measuring periods. A lineshape analysis indicated a small narrowing of the peak atthe beginning of the experiment. This was probably dueto the growth of antiferromagnetic domains when theordered state was being formed.

C. Magnetic-field dependence of the (1 0 0) reflectionfor Bi[011]

To obtain more information about the phase diagramof nuclear-ordered copper, the magnetic-field depen-

Page 46: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

46 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

dence of the antiferromagnetic (1 0 0) Bragg peak wasinvestigated in a series of experiments during which thefinal demagnetization was terminated in different fieldsbelow Bc (Jyrkkio, Huiku, Lounasmaa et al., 1988;Jyrkkio et al., 1989). Special care was exercised to main-tain similar starting conditions in each case before finaldemagnetization to the ordered phases. This was neces-sary since the coil system, shown in Fig. 37, could not beemployed to measure the susceptibility perpendicular tothe external field for determining the initial polarizationusing the procedures described in Secs. IV.A and IV.B.The precooling temperature of the nuclear stages was, ineach case, the same within 10%, and the demagnetiza-tion was always carried out in an identical way. A fur-

FIG. 41. Antiferromagnetic (100) Bragg reflection of copper.The main figure illustrates the integrated neutron count as afunction of time after the zero field had been reached. Theinset shows time dependence of the static susceptibility (in ar-bitrary units). The lower figure (b) depicts the time evolutionof the (1 0 0) peak, as observed by a linear position-sensitivedetector. The four 75-s measuring intervals are centered attimes indicated for each solid curve by the corresponding sym-bol in the upper figure. These bell-shaped curves are the bestGaussian fits to the experimental data. From Jyrkkio, Huiku,Lounasmaa, et al. (1988); Jyrkkio et al. (1989).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

ther check on the experimental starting conditions wasprovided by measurements of the spin-lattice relaxationtime in the paramagnetic regime, when the system hadwarmed above the ordering temperature; the observedt1 was 2062 min in all experiments. On the basis ofcalculations of polarization losses during demagnetiza-tion and using previous susceptibility data (Huiku et al.,1986), it was estimated that, just before entering the or-dered phases, the initial polarization pi50.9660.01, cor-responding to Si50.1R ln4.

The experimental data are presented in Fig. 42. AtB50.04 mT, the qualitative behavior of the neutronsignal was the same as at B50, but the intensity wasless. At B50.08 mT the neutron signal was further re-duced. In all these fields, the static susceptibility x8(0)parallel to the field was about the same, showing for thefirst 4–5 min almost a plateau.

At B50.10 mT, the neutron intensity was close tozero during the entire experiment while, in contrast todata at lower fields, the susceptibility showed a clearincrease for the first 4 min. At B50.12 mT, the neutroncount was drastically different from that at 0.10 mT. Theintensity was very high, as when B50, immediately af-ter the final field was reached, but showed, in contrast tothe low-field behavior, a very rapid decrease at the be-ginning of the experiment. After about 2.5 min no neu-trons were observable above the background. The sus-ceptibility increased almost 20% during the first 4 min.The neutron intensity thus disappeared clearly beforethe susceptibility maximum was reached.

At B50.16 mT, the characteristics were again differ-ent. The neutron intensity was very high initially, as at0.12 mT, but decreased more slowly. The signal disap-peared at the maximum of x8(0). The behavior of thesusceptibility was qualitatively the same as at B50.10mT and 0.12 mT, showing a clear increase in the begin-ning.

At B50.20 mT and 0.24 mT, the neutron signal wassimilar to that at B50, but the intensity was less, espe-cially at 0.24 mT. The increase in susceptibility at B50.16 mT was reduced to a plateau at B50.20 mT,and at B50.24 mT only a short, nonexponential decaywas observed initially. Finally, at B50.30 mT (notshown in Fig. 42), there were no signs of ordering.

The field dependence of the (1 0 0) Bragg reflectionthus shows three distinct regions. In the low- and high-field (close to Bc) regimes, the peak was observed ap-proximately until the susceptibility indicated transitionto the paramagnetic phase. It was therefore concludedthat in these field regions there exist antiferromagneticphases characterized by type-I order.

In intermediate fields, the (1 0 0) signal showed a verylow intensity or a transient behavior, disappearingclearly before the susceptibility indicated paramagnet-ism. It was suggested (Jyrkkio et al., 1989), in fact, thatthe rapidly decaying (1 0 0) signal at B50.12 mT was ametastable trace from the higher-field (1 0 0) phaseformed during demagnetization. It was impossible, how-

Page 47: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

47A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

FIG. 42. Integrated neutron intensity measured for copper at the (1 0 0) Bragg position and the susceptibility x8(0) (insets) asfunctions of time after final demagnetization to the field indicated on each frame; x8(0) is given in arbitrary units. From Jyrkkio,Huiku, Lounasmaa, et al. (1988).

ever, to determine accurately the phase-transition fieldsbetween the intermediate-field phase and the high- andlow-field (1 0 0) phases.

The nature of the intermediate-field phase remainedobscure during the course of the first set of neutron-diffraction experiments. The clear increase of x8(0)pointed to the existence of a well-defined antiferromag-netic phase in this region. However, more exotic inter-pretations, such as the intermediate state’s being an in-commensurate phase or even a spin glass, could not beexcluded.

In fields above B50.16 mT, the neutron data couldbe understood in terms of the picture suggested on thebasis of susceptibility measurements on the single-crystal specimen (see Sec. V.F and Fig. 1) (Huiku et al.,1986). With increasing field, the nuclear spins tilted to-wards B, so that the antiferromagnetic Bragg peak be-came weaker. By extrapolating to the field at which theneutron intensity disappeared, one could obtain a criti-cal field Bc50.25 mT, in agreement with susceptibilitymeasurements.

D. (0 23

23) reflection

An antiferromagnetic Bragg reflection from theintermediate-field phase was searched for during severalexperiments. In the first trials (Jyrkkio et al., 1989), apoint-by-point inspection was made in the reciprocal lat-tice. In the second set of experiments by Annila,Clausen, Lindgard, et al. (1990), it was decided to at-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

tempt scans in the reciprocal lattice to make the searchmore effective. Initially, scanning did not appear feasiblebecause of the inevitable eddy-current heating due tovibrations caused by movements of the diffractometer.It had been anticipated that this would destroy the orderimmediately. Fortunately, in fact, scanning shortenedthe time spent in the ordered state only by about 40%compared with experiments without scanning. This re-duction was large but acceptable since it permittedabout a ten times more efficient search than the station-ary technique.

On the basis of theoretical studies (Lindgard, Wang,and Harmon, 1986; Lindgard, 1988a) an antiferromag-netic Bragg reflection of the form (0 j j) seemed a prom-ising candidate for order in the intermediate-field phase.Owing to the symmetry of the fcc lattice, the (0 j j)point in reciprocal space is equivalent to (1 1-j 1-j).Since the neutron background was lower in the latterdirection, it was chosen for the scans. Figure 43 showsthe result of such a measurement (Annila et al., 1990) inthe field B50.07 mT, which is in the region of the low(1 0 0) intensity. A very clear Bragg peak was found at(1 h h), with h5 1

360.01. The small uncertainty in thecommensuracy resulted from the statistical error of thecounts according to the Poisson distribution and is alsodue to the continuous movement of the spectrometerduring counting. The orientation of the sample was mea-sured with good accuracy afterwards, using the latticereflection (3 1 1) for the l/3 contamination of the neu-tron beam.

Page 48: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

48 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Later, the symmetry-related 6(0 23

23) and (12 1

3213) re-

flections were observed as well. In consecutive experi-ments the neutron intensities of these three signals weremeasured as a function of time at B50.08 mT while thespin system was warming owing to the spin-lattice relax-ation (see Fig. 44). Because the shapes of the warmupcurves were similar, it was concluded that the reflec-tions, indeed, were equivalent.

Next, the neutron intensity of the (1 13

13) peak was

measured at various constant magnetic fields as a func-tion of time while the nuclear-spin system was warmingup after demagnetization. Four examples of warmupcurves are shown in Fig. 45. Neutron counts were re-corded every 0.5 s during the first 150 s. Thereafter datawere taken with a 15-s counting time. The final demag-netization and the subsequent stay at a constant fieldduring warmup are schematically illustrated in the inset.

FIG. 43. (1 13

13) Bragg peak of copper nuclear spins at B

50.07 mT along the @1hh# direction. From Annila et al.(1990).

FIG. 44. Time dependence of the 6(0 23

23) and (1 2

13 2

13)

reflections in B50.08 mT, measured for copper in consecutiveexperiments after similar cooling cycles. The spin temperatureincreases with time. From Annila et al. (1992).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

The data recorded at B50.10 mT (see Fig. 45, curveb) showed a quick increase of the (1 1

313) intensity within

1 s after the termination of the field sweep. The signalthen decreased approximately at a constant rate duringthe next four minutes as the antiferromagnetic sublatticepolarization vanished owing to warmup. Along the tailof the curve, the rate of the intensity reduction becameslower. The change in the shape of the signal may indi-cate a change in the time vs temperature relationship.

As the field strength was reduced, the maximum neu-tron intensity became smaller and its temporal behaviorchanged. At 0.05 mT (see Fig. 45, curve c), the signaldecreased only slowly during the first 2 min. This ten-dency became more apparent when the field was low-ered to about 0.03 mT. A mechanism that tends to coun-teract the decrease in intensity due to the warmup isneeded to explain the observed behavior. Growth of the(0 2

323) domains was suggested as an interpretation (An-

nila et al., 1990).In very low fields, below 0.03 mT, the (1 1

313) intensity

quickly disappeared. However, even at B50.01 mT(see Fig. 45, curve d) some intensity was observed ini-tially, probably caused by the remaining spin order thathad been formed during demagnetization through thehigher-field regions.

The maximum (1 13

13) intensity decreased rapidly in

fields above 0.10 mT. At B50.13 mT (see Fig. 45, curvea) a weak signal built up in 30 s and disappeared quitesoon, in about 1.5 min. A maximum of about 10 counts/swas observed. Obviously, only a small portion of thecrystal had developed long-range (0 2

323) order. This

phase was not stable in high magnetic fields where the(1 0 0) reflection was strong.

FIG. 45. Time dependence of the (1 13

13) reflection for copper

at (a) B50.13 mT, (b) 0.10 mT, (c) 0.05 mT, and (d) 0.01 mT.A schematic phase diagram in the B-T plane is shown by theinset, where arrows indicate demagnetization into the orderedphase along a vertical isentrope and subsequent measurementsduring horizontal warmup in the above-mentioned constantmagnetic fields. From Annila et al., (1992).

Page 49: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

49A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

E. Neutron-intensity contour diagram

A concise presentation of the data from 18 experi-ments was given in Fig. 3 in the form of a neutron-intensity contour diagram. In this graph, new data on the(1 0 0) peak, measured in the same way as the (1 1

313)

reflection, were supplemented with the earlier results ofJyrkkio and co-workers (Jyrkkio, Huiku, Lounasmaaet al., 1988; Jyrkkio et al., 1989). The most important dif-ference between the two sets of experiments is that thefinal demagnetization had been made at the sweep rateof 10 mT/s in the earlier measurements, whereas dB/dt560 mT/s was used in the later experiments by Annilaet al. (1990, 1992). Previously it had been observed thatthe faster the field sweep was in the high-field region,the more (1 0 0) intensity was detected during the sweep(Siemensmeyer et al., 1990). This effect could result ifthe (1 0 0) domains were preferred over the (0 1 0) and(0 0 1) domains. No such sweep-rate dependence wasfound for the (1 1

313) intensity.

Three maxima occur in the contour diagram: at B50.09 mT for the (1 1

313) peak, and at B50 and B

50.15 mT for the (1 0 0) reflection. The (1 13

13) signal is

strongest when the (1 0 0) signal is weakest and viceversa, suggesting the presence of three distinct phases.The upper boundary is sharp, with the contours closelyspaced, whereas in low-intermediate fields, 0.02<B<0.06 mT, both reflections coexist over a wide region.

F. Kinetics of phase transitions when Bi[011]

Several features of the neutron-diffraction data can beassociated with the decay and growth of nuclear mag-netic order. The time scale of the kinetics is long com-pared with that of electronic magnets because of theweak interactions between nuclear spins. In this respectnuclear magnets are very suitable systems for directstudies of the ordering processes. On the other hand,conditions for true thermal equilibrium cannot be pro-duced readily because of the long time constants.

1. Initial temporal development of neutron intensity

The temporal behaviors of the neutron signals in thetwo transition regions were different (see Fig. 46). AtB50.12 mT, the (1 0 0) intensity disappeared while the(1 1

313) signal appeared about 30 s after the demagnetiza-

tion had ended. The background level was reached in afew minutes. The transients show how the (1 0 0) struc-ture, formed during the field sweep through the upperphase, transforms into the (0 2

323) spin configuration. The

transition is sharp, since the width of the phase bound-ary appears to be on the order of the field inhomogene-ity DB50.01 mT.

In the lower 0.04-mT transition region, the signals didnot show the initial kinetics and decreased only gradu-ally during the first 1.2 min, whereafter the typical tail ofa warmup curve was observed. Since no transients wereseen and both signals persisted for a long time, withsimilar temporal behaviors, it must be concluded that inthis region, 0.02 <B<0.05 mT, the (1 0 0) and the

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

(0 23

23) orders coexist, either in the form of domains of

two different spin configurations or as a combined(1 0 0) and (0 2

323) multiple-k structure. In the former

case, the (1 0 0) and (0 23

23) configurations must be nearly

degenerate in energy. In the latter, a combination struc-ture continuously transforms from the pure (0 2

323) order

to the (1 0 0) order when the magnetic field is decreased,and the lower transition is not a phase transition at all.

2. Hysteresis at phase boundaries

The phase boundaries were probed by successivelysweeping the magnetic field down and up across thetransition regions [see Fig. 47(a), inset]. At the upperboundary around B50.12 mT [see Fig. 47(b)], when thefield was lowered, the (1 0 0) signal disappeared at fieldssmaller than those at which it reappeared when the fieldwas raised again. Accordingly, the (1 1

313) counts began

to increase at smaller fields when B was swept down-ward as compared with the field where counts started todecrease when the field was being swept upward. In thisway a pair of hysteresis loops were created.

In the low-field region, loops were observed as well,but the changes in the intensities were opposite to thoseat the upper boundary [see Fig. 47(c)]. Hysteresis wasseen over the same field interval where the (1 0 0) and(1 1

313) counts had similar time evolutions. The two tran-

sition regions appeared similar during the field-sweepexperiments, in contrast to the constant-field measure-ments (see Fig. 46).

The boundary across Bc(T) to the paramagneticphase was examined in a similar manner for comparison[see Fig. 47(a)]. No hysteresis was found in this case,consistent with a second-order phase transition.

FIG. 46. Neutron counts from copper nuclear spins vs timenear the upper phase boundary at B50.12 mT and in thelower transition region at B50.04 mT: s , (1 0 0) peak; d ,(1 1

313) reflection. From Annila et al. (1992).

Page 50: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

50 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

FIG. 47. Neutron counts for copper vs magnetic field (a) near Bc and at (b) the upper, and (c) the lower transition regions whenthe magnetic field was swept across the phase boundaries at the rate 8 mT/s [see inset in (a)]. s , (1 0 0) peak; d , (1 1

313) reflection.

From Annila et al. (1992).

3. Temporal changes in the width of the Bragg peak

Figure 48 shows the peak width of the (1 0 0) Braggreflection during warmup in three different externalfields. The error bars are those given by the fitting rou-tine, and they provide an estimate on the relative qualityof the fit. The data show that for all applied fields a smallnarrowing is observed at first. This can be explained bythe growth of the antiferromagnetic domains. The peakwidth is smallest in fields close to B50 and broader inhigher fields; the broadest peak was observed at B50.12 mT.

Separate measurements on samples of different mate-rials, such as a germanium crystal and Al 2O 3 powder,showed that the instrumental resolution varied from 0.8°to 1.1°. For example, the first data points at B50.20 mTand all the B50.12 mT data are clearly broader thanthe instrumental peak width. Since B50.12 mT is in the

FIG. 48. Widths of the (1 0 0) Bragg peaks in copper as func-tions of time, deduced from Gaussian fits to the linear-detectordata in three selected fields. From Jyrkkio et al. (1989).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

region of the upper phase boundary, the larger peakwidth is not surprising. The system is probably in ametastable state and does not find its way to a well-defined state during the short lifetime of the neutroncount. A rough calculation on the broadening of thepeak in the beginning of the B50.12 mT experimentindicates domain sizes on the order of 500 Å.

4. Decay of metastable states

Relaxation times associated with magnetic ordering inthe low-field region were also investigated in the courseof the two experiments illustrated in Fig. 49. The timeevolution of the neutron intensity was measured in fieldsB510.04 mT and B5 –0.04 mT, where the negativesign indicates that, during the final demagnetization, thefield was swept through and beyond zero. The character-istic difference was the very fast decay of the 0.04 mTsignal during the first 1.5 min. It was believed that this

FIG. 49. Time evolution of the (1 0 0) neutron intensity forcopper in two successive experiments: s , demagnetizationstopped at B50.04 mT; d , demagnetization continuedthrough zero to the same field in the opposite direction. FromJyrkkio et al. (1989).

Page 51: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

51A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

was again due to metastability, i.e., a remanent, quicklydisappearing trace of the spin arrangement created closeto B50. After 1.5 min, the two signals were behaving inthe same way, which indicates symmetry in the phasediagram around B50.

Relaxation times in higher fields showed interestingfeatures as well. The neutron count rate at the (1 0 0)reflection in fields from 0.1 to 0.17 mT, immediately af-ter the end of demagnetization, was found to depend onhow quickly the field had been decreased to this region.Intensities measured for three different sweep rates arepresented in Fig. 50 (Siemensmeyer et al., 1990). Duringthe fastest scans the intensity reached a value almostthree times higher than what was observed in the experi-ments under a constant field with a counting time of 15s. There must be a fast relaxation process that causes thequick decay of the (1 0 0) intensity. From the sweeprates the relaxation time of this process was estimated tobe in the range between 2 and 5 s.

Intriguing behavior was observed when the fieldsweep with rate 211 mT/s was continued through zeroto negative fields. First, only a weak (1 0 0) reflectionwas observed in the high-field phase, when B520.12 – 20.25 mT, as is shown by the data points in-dicated by rectangles in Fig. 50. The field was changeduntil B520.3 mT was reached, at which point the di-rection of the sweep was reversed. The spins were thenin the paramagnetic phase since Bc50.25 mT. The an-tiferromagnetic phase was reentered at the sweep rate of11 mT/s, as shown by the neutron data illustrated bysolid circles. The (1 0 0) reflection in the high-fieldphase, B520.12 – 20.25 mT, was approximatelytwice as high as the intensity observed during the firstfield sweep. This was surprising since antiferromagneticorder should decay with time due to the spin-lattice re-

FIG. 50. Neutron intensity at the (1 0 0) Bragg position duringfield changes at three different rates. The measurements withnegative sweep rates proceeded from a positive to a negativefield. From Siemensmeyer et al. (1990).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

laxation process. A possible explanation is that duringthe first field sweep, when the (1 0 0) intensity was low,spins partly remained in the metastable state with(0 2

323 ) order.

5. Entropy losses

Nonadiabatic behavior at the phase boundaries wasfurther investigated by Annila et al. (1992) in a series ofexperiments. The magnetic field was first reduced to aselected value Bi above, in, or below the (0 2

323) state,

then raised well above Bc , and finally lowered to Bf50.18 mT, where the (1 0 0) neutron counts were re-corded (see Fig. 51, inset). This elaborate sequence offield sweeps assured that the measuring field Bf was al-ways reached at the same time and in the same way,independently of Bi . Therefore the spin structure cre-ated at Bi could not possibly affect the distribution ofthe total intensity among the (1 0 0), (0 1 0), and (0 0 1)domains and thus the measured (1 0 0) neutron inten-sity. Counts collected at Bf were summed in order toacquire better statistics, and all the data were scaled by areference measurement at Bi50.12 mT (see Fig. 51).

It was found that the total neutron count at Bf clearlydropped when Bi was reduced from 0.12 mT to 0.08 mT.The midpoint and the width of the decrease coincidedwith the center and the width, respectively, of the hys-teresis loops along the upper phase boundary [see Fig.47(b)]. The loss of intensity was related to a gain ofentropy during a first-order phase transition. This alsoexplains, at least partly, why the (1 0 0) intensity at zerofield was lower than at B50.16 mT (see Sec. VII.C);part of the signal was simply lost, owing to nonadiaba-ticity, when the upper phase boundary was crossed. Thenature of the lower transition could not be resolvedwithin the experimental error.

G. Comparison with simultaneoussusceptibility measurements

The longitudinal, low-frequency susceptibility xL wasmeasured simultaneously with the neutron signal from

FIG. 51. Total (1 0 0) neutron counts from copper nuclearspins measured at Bf50.18 mT vs the magnetic field Bi . Thesample was brought to Bf from the paramagnetic phase after ashort stay in Bi (see inset). From Annila et al. (1992).

Page 52: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

52 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

the lower section of the sample (see Fig. 37). A contourmap (see Fig. 52), similar to the neutron-intensity dia-gram (see Fig. 3), was composed from measurements atconstant fields (Annila et al., 1992). A comparison of thetwo graphs helps to determine whether the antiferro-magnetic behavior of the susceptibility signal can be at-tributed to the (1 0 0) and (0 2

323) reflections only, or

whether other Bragg peaks should be present as well.The critical field is about 0.2660.01 mT according to

the xL data, compared to Bc50.25 mT obtained byneutrons. The phase boundary of the ordered state,determined from xL , closely engulfs the outermostneutron-intensity lines of Fig. 3. The clearest discrep-ancy occurs around B50.13 mT. The outer contours ofthe (1 0 0) and (1 1

313) signals bend strongly inwards be-

fore crossing, which might indicate a transition into thedisordered state only 1.5 min after demagnetization andhence may point to a re-entrant Bc(T) curve around thisfield. However, the spins seem to be antiferromagneti-cally ordered for t.1.5 min as well, because the maxi-mum of xL is observed at t52.5 min. This suggests an-other phase, with neither (1 0 0) nor (0 2

323) modulation

in the time interval t51.5 –2.5 min.The clear minimum in xL at intermediate fields (see

Fig. 52) coincides in field and time with the maximum ofthe (1 1

313) intensity. The neutron maxima of the (1 0 0)

reflections (see Fig. 3) do not correspond to distinctivefeatures in the susceptibility diagram. The phase transi-tions also look different in the two graphs. The sharpupper boundary in the neutron diagram is contrastedwith a smooth decrease of susceptibility. In low fields, inthe coexistence region of the slowly changing (1 1

313) and

(1 0 0) intensities, there is a local maximum of xL at B50.04 mT. The differences between the susceptibility

FIG. 52. Longitudinal susceptibility of copper shown by con-tours in arbitrary units as a function of time and magnetic fieldalong the [01 1] direction. The antiferromagnetic region (AF) isbounded by the paramagnetic phase (P) where the contourseventually become diagonal and uniformly spaced. The whitecurve follows the outermost neutron-count-rate contours ofFig. 3. The antiferromagnetic region extends slightly beyondthis curve; the difference can be especially large in fieldsaround 0.13 mT owing to the presence of the (6

23 6

23 0) and

(623 0 6

23) signals shown in Fig. 57(c). These reflections have a

different field dependence than the (0 23

23) order illustrated in

Fig. 3. Modified from Annila et al. (1992).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

and the neutron-diffraction data in the appearances ofthe transition regions might be better understood if thetwo other Cartesian components of susceptibility hadbeen measured, as was done earlier for Bi[001] (Huikuet al., 1986).

Following this reasoning, discussed in Sec. V.F in thecontext of Fig. 33, Annila et al. (1992) concluded, on thebasis of their xL data, that in high fields the antiferro-magnetic sublattice magnetization of the (1 0 0) struc-ture is perpendicular to B. When the magnetic field islowered, the sublattice magnetization becomes moreand more parallel to the field and reaches its extremalorientation at B50.09 mT, corresponding to the (1 1

313)

reflection. Further reductions of the field causes thesublattice magnetization to become more perpendicularto the field. The extremum is at about B50.04 mT. Instill lower fields, part of the sublattice magnetization ap-pears to be parallel to the field as well.

H. Other field directions

An obvious extension to the neutron-diffraction ex-periments on copper described so far was to examine thephase diagram with the magnetic field aligned alongcrystallographic axes other than the [01 1] direction. Thisis important because not only the spin configurations butalso the number of phases, as predicted by theoreticalstudies, might differ for various field directions. The re-sults of these experiments have been described by An-nila et al. (1992).

The external field B could be applied along an arbi-trary direction by using an orthogonal pair of saddlecoils and a split-pair solenoid, perpendicular to thesaddle coils (see Fig. 53). The magnetic structures couldthen be studied by measuring the (1 0 0) and (0 2

323) re-

flections in various orientations of the field. Figure 54illustrates how the high-symmetry directions of the firstBrillouin zone are related to the scattering plane.

Because it would have been impractical to investigatethe phase diagram in the same detail as was done for the[01 1] field direction by means of a series of warmupmeasurements in a constant field, a new scheme was em-ployed to establish first an overall picture. Furthermore,it was technically difficult to fix the field accurately alongall desired axes for a long time, since the demagnetiza-tion field (see Sec. V.G) kept decreasing owing to spin-lattice relaxation. Therefore the field was repeatedlyswept from B50.3 mT.Bc down to zero and up again,with dB/dt510 mT/s. At this rate, one cycle took 60 sand a reduction of polarization during that period didnot affect the field in the specimen substantially, even inthe worst cases.

Annila et al. (1992) took great care to maintain iden-tical initial conditions for the field-sweep experiments.Reproducibility of the data was not usually investigatedto save beam time.

1. [100] directions

The three directions [100], [010], and [001] are equiva-lent under cubic symmetry. The latter two are symmetric

Page 53: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

53A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

with respect to the accessible (1 0 0) and 6(0 23

23) posi-

tions in the scattering plane (see Fig. 54). Therefore itsufficed to make measurements at Bi[100] and [001].Data for the first down-and-up sweeps are shown inFig. 55.

FIG. 53. Schematic view of the coil system surrounding thesample inside the 4.6-T split-pair superconducting solenoid.The crystallographic [100], [011], and [01 1] directions areshown with respect to the x , y , and z axes of the specimen.Components of the external magnetic field along the Cartesianaxes are generated by three field coils, Bx , By , and Bz . Theupper part of an astatically wound small detection coil formeasuring the susceptibility surrounds the lower end of thesample. From Annila et al. (1992).

FIG. 54. The star of k: the first Brillouin zone of an fcc latticewith some of the crystallographic directions shown. Under cu-bic symmetry, there are three equivalent (1 0 0) positions(s) and 12 equivalent (0 2

323) positions (d). In the first Bril-

louin zone, only the (1 0 0) and the 6(0 23

23) points were in the

scattering plane (shaded) and accessible to the experiments.From Annila et al. (1992).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

The signals were clearly different in the two direc-tions. In high fields, the (1 0 0) intensity was very low forBi[100] whereas it was high in the [001] field direction.In low fields, the (1 0 0) reflection was observed for bothdirections, but its intensity in the [001] direction washigher by a factor of 3. The (0 2

323) reflection was not

present at all for Bi[100], but with Bi[001] the (0 23

23)

signal was observed at intermediate and low fields.From Fig. 55(b) one obtains a critical field of about

0.27 mT and a phase transition to the intermediate-fieldstructure at B'0.13 mT. These fields at which a phasechange occurs are the same as for Bi[01 1] (Fig. 47)within the uncertainty in B .

When B50, one would expect equal intensities, asthe field no longer breaks the symmetry. Because theobserved signals clearly differed (see Fig. 55), this mat-ter was further investigated by two experiments in whichB was rapidly swept to zero along the [100] and [001]directions, respectively, whereafter neutrons werecounted as a function of time (see Fig. 56).

The signals were quite different for the two directionsof the demagnetization field. The data demonstrate thatthe magnetic structures were created during demagneti-zation and determine how the order was divided amongthe (1 0 0) magnetic domains when B50. The time scalerequired for thermal equilibrium between domains islong compared to the measuring time available.

FIG. 55. Neutron intensity of the (1 0 0) (s) and (0 23

23) (d)

peaks for copper vs the magnetic field for downward sweepsalong (a) the [100] and (b) the [001] field directions at the rate10 mT/s. The counts collected when the magnetic field wasraised are marked by ( for the (1 0 0) peak and by ; for the(0 2

323) peak. The directions of the field sweeps are indicated by

arrows as well. From Annila et al. (1992).

Page 54: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

54 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

2. [110] directions

For the [110] field alignments, there are two non-equivalent directions with respect to the (1 0 0) reflec-tion and three directions for the (0 2

323) peak (see Fig.

54). The results from field sweeps along the [011], [01 1],and [1 01] directions are shown in Fig. 57. Even thoughthe [01 1] alignment had been investigated in a largenumber of experiments, as described in the previous sec-tions, field sweeps in this direction were included as wellto facilitate direct comparison with the data in the otherdirections.

In high and low fields, the (1 0 0) signal was observedin all three directions. For the (1 0 0) peak, the [011] and[01 1] directions are equivalent under fcc symmetry;equal neutron signals thus ought to be recorded [seeFigs. 57(a) and (b)]. Experimentally, this was true onlyto the extent that the (1 0 0) reflection was seen over thesame field region for both directions, but the intensitieswere unequal. A small misalignment of the external fieldor slightly different cooling conditions, affecting the ini-tial polarizations and warmup rates during the two ex-periments, could be responsible for this effect. Dispro-portionate signals between the high- and low-fieldregions for the [011] and [01 1] directions may have re-sulted from unbalanced populations of the various do-mains.

When sweeping up again from B50, the high-field(1 0 0) signal was much smaller than expected. Even ifthe intensity loss due to entropy gain during the first-order transition (see Fig. 51) and the subsequent war-mup were taken into account, according to the data inFig. 3, it was expected that at least one-fourth of theinitial counts should have been recovered. Moreover,the susceptibility signal clearly showed that the systemwas not in the immediate vicinity of a transition to theparamagnetic phase. This finding is in accordance withearlier observations (Siemensmeyer et al., 1990) de-scribed in Sec. VII.F.4 (see also the data in Fig. 50).

FIG. 56. (1 0 0) neutron intensity for copper as a function oftime after demagnetization to B50 along the [001] and the[100] field directions at the rate 50 mT/s. From Annila et al.(1992).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

In intermediate fields, the three (0 23

23) signals of Fig.

57 are different. When Bi[011] the reflection was absent[see Fig. 57(a)], whereas for Bi[01 1] [see Fig. 57(b)] andBi[1 01] [Fig. 57(c)] the (0 2

323) reflection was there, but

the signals were different both in intensity and in theirmagnetic-field dependence. The (0 2

323) order was found

in higher fields for Bi[1 01] than for Bi[01 1]. Annila et al.(1992) took this as evidence for two different spin struc-tures with (0 2

323) order.

FIG. 57. Neutron intensity of the (1 0 0) (s) and (0 23

23) (d)

peaks for copper vs the magnetic field for downward sweeps atthe rate 10 mT/s with B along the (a) [011], (b) [01 1], (c) [1 01],and (d) [101] directions. The counts measured during upwardfield sweeps are marked by ( for the (1 0 0) peak and by ; forthe (0 2

323) peak. From Annila et al. (1992).

Page 55: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

55A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

The new (0 23

23) phase observed for Bi[1 01] provides a

possible explanation, as well, for the sweep-rate-dependent (1 0 0) intensity, discussed in Sec. VII.F.4. Asis shown by Fig. 57(c), the (0 2

323) intensity starts to in-

crease already around B50.17 mT for dB/dt510 mT/s,which is approximately the field below which the sweep-rate-dependent (1 0 0) intensity was observed. Takinginto account that there are altogether eight equivalent(0 2

323) reflections with respect to the [1 01] field direc-

tion, one can estimate that at B50.12 mT, for example,the total intensity of these reflections is 2–3 times higherthan the intensity of the (1 0 0) peak. Therefore a char-acteristic time scale, on the order of a few seconds, forthe growth of the (0 2

323) order would explain the sweep-

rate effect.Annila et al. (1992) suggested, on the basis of the hys-

teresis loops in Fig. 57(c), that the low- and high-fieldboundaries of the (0 2

323) phase for Bi[1 01] are of first

order.Fig. 57(d) shows the neutron intensity of the (0 2

323)

reflection when the magnetic field was swept to zeroalong the [101] direction. The shape of the signal is simi-lar to that of the (0 2

323) reflection in Fig. 57(c), as it

should be because these two crystalline directions areequivalent with respect to the (0 2

323) position.

3. [111] directions

For a field aligned along any of the [111] axes, thethree (1 0 0) peaks are equivalent. For the (0 2

323) reflec-

tion, there are two nonequivalent directions (see Fig.54), for example, [111] and [11 1], which were chosen formeasurements.

The most striking feature of the data in Fig. 58 is themissing high-field (1 0 0) phase. This perplexing questionis discussed separately in Secs. VII.J. and XV.G. Fromthe susceptibility measurements the critical field Bc50.26 mT was deduced.

The (0 23

23) peak was absent when B was parallel to

[111], but it was observed at intermediate fields for the[11 1] alignment. The transition from the high-field phaseto the (0 2

323) structure took place at B50.13 mT. The

form of the (0 23

23) signal during the field sweeps was

rather similar to that with Bi[01 1]. At low fields the(1 0 0) peak appeared but with somewhat different in-tensities for the two field directions. The region withoverlapping (1 0 0) and (0 2

323) reflections resembled the

low-field region when Bi[01 1].

4. Selection of stable domains by external-field alignment

The field-sweep experiments illustrated in Figs. 55, 57,and 58 contain a great deal of information on both thestatic and the dynamic properties of the nuclear-spinsystem in copper. As for the equilibrium spin structure,it is important to know how the total order is dividedamong the different domains which, without thesymmetry-breaking external field, would be equivalentin a cubic crystal. To this end, it is useful to show theessentials of the field-sweep data in a different form. InFig. 59, the neutron intensities of the three (1 0 0) and

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

the twelve (0 23

23) reflections have been presented using

a gray scale. Rather than illustrating the results for thetwo fixed reflections with different alignments of B, thedata are displayed by fixing the direction of B with thecrystalline axes and rotating the reflections accordingly.The field-sweep data towards B50, in Figs. 55, 57, and58, have been employed in constructing the diagram.The measured signal of a symmetry-related reflectionwas assigned to those (1 0 0) and (0 2

323) reflections

which were not directly observed. When results fromequivalent field directions were available, data from theexperiment for which the demagnetization correctionwas smallest (see Sec. V.G) were used.

The data emphasize the fact that an application of thefield completely destroys the cubic symmetry in thesense that definite selection rules are observed. For ex-ample, in the high-field phase when B was parallel to[100], no neutron signal was seen for the (1 0 0) Braggreflection although there was intensity at the (0 1 0) and(0 0 1) positions. In the intermediate-field region, someof the twelve (0 2

323) reflections had strictly zero inten-

sity, while other positions showed a high intensity. Theseselection rules provide important tests for theoretically

FIG. 58. Neutron intensity of the (1 0 0) (s) and (0 23

23)

(d) peaks for copper vs the magnetic field for downwardsweeps at the rate 10 mT/s with B along the [111] (upperframe) and [11 1] (lower frame) directions. The counts for up-ward sweeps are marked by ( for the (1 0 0) peak and by ;for the (0 2

323) peak. From Annila et al. (1992).

Page 56: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

56 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

FIG. 59. Neutron intensity, according to the gray-scale shown below, for the three (1 0 0) and the twelve (0 23

23) Bragg reflections

from copper nuclear spins as a function of the magnetic field parallel to the three high-symmetry crystallographic directions. FromAnnila et al. (1992).

calculated spin structures (see Sec. XV.F.2). The experi-mental data are in agreement with the spin configura-tions predicted earlier by Viertio and Oja (1987, 1990b).

The selection rules observed especially for the (0 23

23)

reflections in copper differ in important respects fromtypical domain selections by the field, as observed inelectronic magnets. Usually, application of B stabilizesthose zero-field domains which display sublattice mag-netization perpendicular to the field, while domains withsublattice magnetization along the field become un-stable. In copper, however, (0 2

323) is not stable at B50

to begin with. Instead, the field induces order for certainBragg peaks among the twelve cubic-symmetry-related(0 2

323) reflections.

Annila et al. (1992) were able to draw some conclu-sions on the spin configurations based on the data pre-sented in Fig. 59. They reasoned as follows.

For Bi[100] in high fields, 0.10,B,0.25 mT, the cop-per sample contained either two types of domains, inwhich the antiferromagnetic order propagates along k =(p/a)(0,1,0) or k5(p/a)(0,0,1), respectively, or therewas a simultaneous modulation along both vectors. Inthe former case the structure would be single-k and inthe latter double-k, but a distinction could not be madeon the basis of the experimental data. In intermediatefields, B,0.12 mT, the (0 2

323) modulations were not ob-

served for k'B, but they were clearly seen for k makinga 45° angle with B. Again it was not possible to distin-guish between multidomain and multiple-k phases.

For Bi[011] in high fields, the (1 0 0) order was ob-served in addition to the two symmetrically equivalent(010) and (001) propagations. The field dependencieswere somewhat different. One can conclude that thespin configuration was not a single-k structure, sincethen only the (1 0 0) reflection or, alternatively, only the(0 1 0) and (0 0 1) Bragg peaks would have been ob-served. The structure could be a triple-k state or adouble-k spin configuration composed of (1 0 0) and

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

(0 1 0) propagations or, equivalently, of (1 0 0) and(0 0 1) propagations.

In intermediate fields, those (0 23

23)-type reflections

for which kiB were absent, those with k'B were mostintense, and those with k making a 60° angle with Bwere strong, too. There are most certainly two different(0 2

323)-type structures, because the magnetic-field de-

pendence for k'B was clearly different from that for theother eight propagations.9

For Bi[111] in high fields, the ordered phase was notidentified. In intermediate fields, the six (0 2

323) modula-

tions perpendicular to B were observed.At least in zero field, the domain population is af-

fected by the route along which the spins are demagne-tized to B50, as was clearly shown by the data in Fig.56. This raises the question of to what extent the sameapplies in higher fields. The experimental, clear-cut se-lection rules for the (0 2

323) and the (1 0 0) reflections

suggest, however, that equilibrium domain populationswere observed in intermediate and high fields. Besides,when the system enters the ordered region at B5Bc50.26 mT, there is no antiferromagnetic order to startwith, and the developing domain population shouldtherefore correspond to the equilibrium spin distribu-tion.

I. Intensity diagram for different field alignments

The ambitious goal of the work by Annila et al. (1992)was to construct the phase diagram for different direc-tions of the external magnetic field in the plane contain-ing the high-symmetry crystalline alignments [100],[011], and [111]. In order to span the directions between

9In their first paper, Annila et al. (1990) indirectly but erro-neously concluded that only k ' B reflections were present.

Page 57: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

57A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

these axes, Annila et al. supplemented the results of thefield sweep measurements (see Sec. VII.H) with datafrom experiments during which the magnetic field wasrotated at a constant strength. These field-rotation mea-surements were done at B50.09 mT and at B50.16mT, for the (0 2

323) and (1 0 0) reflections, respectively.

The two neutron peaks are most intense in these fields.The results of one rotation experiment are illustrated

in Fig. 60. The external field was turned first from the[100] axis, via the [11 1] field alignment, parallel to the[01 1] direction. A reverse rotation was then performed.Both field rotations took 30 s. At B50.09 mT, the(0 2

323) intensity smoothly increased as the field was

turned. Apart from some hysteresis, the reverse rotationdisplayed rather similar behavior. The absence of anylarge-intensity loss suggests that, during rotation, thereis only one complicated multiple-k structure in thesample or several different structures separated by con-tinuous transitions.

At B50.16 mT, the (1 0 0) signal was initially slightlyabove the background but soon disappeared as the rota-tion continued. At the [01 1] field alignment, no neutronsabove the background were observed, in disagreementwith a large number of measurements that had been per-formed earlier in this field direction (see Sec. VII.C).The spins were then momentarily brought into the para-magnetic phase by increasing the field above Bc and byimmediately lowering it back to 0.16 mT. A clear (1 0 0)reflection was then observed. When the field was rotatedback towards the [100] alignment, the signal vanishedbefore the [11 1] direction was reached, consistent withthe field-sweep measurements of Fig. 58.

Supplemented with the field-rotation data, a neutron-intensity diagram was constructed on the basis of the

FIG. 60. Neutron intensity for the (1 0 0) peak (s) at B50.16 mT and for the (0 2

323) reflection (d) at 0.09 mT when

the magnetic field on the copper specimen was rotated fromthe [100] axis, u50°, to the [01 1] direction, u590°, at a rateof 3°/s. The counts collected during reverse turns are markedby ( for the (1 0 0) peak and by ; for the (0 2

323) peak.

Directions of the field rotations are indicated by arrows aswell. From Annila et al. (1992).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

field sweeps to B50. In Fig. 61, the neutron counts ob-tained from reflections with the same symmetry with re-spect to the field direction were added together. Theintensities measured for the symmetric positions wereassigned to those Bragg peaks which were not directlyobserved. Ideally this would be equivalent to observingall three (1 0 0) and all twelve (0 2

323) reflections. Since

the experiments were affected by time-dependent phe-nomena, hysteresis, and entropy gains, the neutron-count contours are approximate but, nevertheless, givean overall picture of the ordered phases.

The antiferromagnetic state in Fig. 61 is bounded bythe critical-field line Bc , which is presumably of secondorder for all directions. Bc was determined from theneutron data and susceptibility measurements. In highfields, there are two (1 0 0) phases. One resides over awide span of directions around Bi[100] and the other isover a smaller region about Bi[011].

The ordering vector for the high-field phase forBi[111], is a puzzle. The region with no neutron intensitycovers a large area of the phase diagram (see Fig. 61).The unknown structure seems to be metastable in the[011] directions as well, because, after the field belowBc was rotated via the [11 1] alignment parallel to the[01 1] axis, the (1 0 0) signal remained absent but couldbe recovered after a sweep to a field above Bc and back(see Fig. 60).

The transitions from the high-field phases to the(0 2

323) structures are all presumably of first order. A

pure (0 23

23) phase is present only over a narrow interval

in all directions, before the low-field (1 0 0) phase begins

FIG. 61. Neutron intensity of nuclear spins in copper as a func-tion of the magnetic field in the plane of the three high-symmetry directions. Counts for field alignments equivalentunder the fcc symmetry are summed together. Contours(neutrons/s) for the (1 0 0) intensity are red and for the(0 2

323) intensity they are green. The antiferromagnetic phase is

bordered by the second-order Bc curve. Modified from Annilaet al. (1992).

Page 58: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

58 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

FIG. 132. Low-temperature records: greentriangles, superfluid 3He; green inverted tri-angles, solid 3He; green diamond, 3He/ 4Hemixture; red spheres, conduction electrons incopper; yellow sphere, Bose-Einstein conden-sation in gaseous Rb; white squares, nuclearspins in Cu, Ag, or Rh; white triangles, nega-tive nuclear-spin temperatures in Ag or Rh.For references, see Sec. XVI.C.

to emerge. Therefore the same pattern as was alreadyobserved for Bi[01 1], namely, a (0 2

323) order coexisting

with a (1 0 0) structure over a relatively wide interval inlow fields, seems to be valid for all directions. The zero-field ground state apparently is of pure (1 0 0) type.

J. Search for other antiferromagnetic Bragg peaks

Since the (1 0 0) reflection was not observed in thehigh-field region with B along a [111] crystalline axis,several experiments were performed to look for otherantiferromagnetic Bragg peaks (Annila et al., 1992).Also, during some earlier stages of the Risø measure-ments, the reciprocal lattice was searched for new reflec-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

tions (Jyrkkio et al., 1989; Annila et al., 1990). No neu-trons above the background were observed in any ofthese experiments.

Figure 62 summarizes the positions in the reciprocallattice that have been investigated when the externalfield was applied along the [01 1] direction. Thesearch scans, which are marked by thick lines, were per-formed at B50.07 mT and also at B50 because the(1 0 0) signal in zero field appeared to be too small incomparison with its intensity at fields around B50.16mT. Several commensurate positions, marked by solidsquares, were also investigated. The possibility of a re-flection at (0 1

313), which could be a harmonic component

of the (0 23

23) peak, was inspected over the whole field

range below Bc , but no neutrons above the backgroundwere found. Several other commensurate points were

Page 59: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

59A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

investigated as well, but with the same negative result.For a field applied in the [111] direction, the (0 1

212)

and ( 12

12

12) positions, which correspond to conventional

type-II and type-IV antiferromagnetic structures (Smart,1966; see also Fig. 110), were investigated in all fieldsbelow Bc but again with negative results (Annila et al.,1992). Type-III reflections were not in the scatteringplane and thus could not be examined. The scans at B50.17 mT for Bi[11 1] covered the (0 h h) line in thereciprocal lattice from h5 1

3 to h5 34 (the K point) and

the (h 0 0) line from h5 13 to h51 (see Fig. 62). Disap-

pointingly, no neutrons above the background were de-tected. A similar scan along the (0 h h) line should havebeen performed also, using the Bi[111] field alignment,as this would have tested one of the subsequent theoreti-cal suggestions for the ordering vector (Oja and Viertio,1992).

Theoretical calculations of the spin structure in thehigh-field phase for Bi[111] will be discussed in Sec.XV.G.

VIII. SUSCEPTIBILITY AND NMR DATA ON SILVER

An introduction and overview of investigations onnuclear magnetism in silver was given in Sec. I.E. Thespin Hamiltonian can be written in the formH5HD1HRK1HZ . The dominating spin-spin energy isthe nearest-neighbor antiferromagnetic exchange inter-action HRK , proposed by Ruderman and Kittel (1954).The dipolar force HD between nearest neighbors issmaller by a factor of 3 in silver. This is the most impor-tant difference between silver and copper; HD for neigh-boring copper nuclei is larger than HRK by a factor of 2.Owing to the strong exchange interaction, the spin sys-tem in silver bears a close resemblance to an fcc Heisen-

FIG. 62. First Brillouin zone of the fcc lattice in the [01 1]scattering plane. The search scans are marked with heavy lines:d , the observed 6(0 2

323), (1 1

313), and (1− 1

3− 13) reflections;

s , the observed (1 0 0) reflections; j , commensurate posi-tions, where neutron intensity was not found. From Annilaet al. (1990).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

berg antiferromagnet and has been the object of muchtheoretical interest; the ground-state properties are af-fected by frustration (Binder and Young, 1986). Becausethe nuclear spin I5 1

2, quantum effects are expected tobe prominent.

Technically, the essential difference between nuclear-ordering studies of silver and copper is the smaller mag-netic moment of silver nuclei, which results, e.g., inweaker spin-spin interactions, a lower ordering tempera-ture, and weaker coupling to the conduction electrons.These features made it possible for the Helsinki groupto extend earlier studies of nuclear magnetism in metalsinto a new regime, namely, to negative spin tempera-tures. The general features of nuclear magnetism atT,0 were discussed in Sec. II.D.

A. Introduction to experimentsat negative spin temperatures

Purcell and Pound (1951) first produced negative tem-peratures by means of population inversion, using LiF asthe working substance. The implications of these earlyNMR experiments, in which t1'5 min and T'21 Kwere reached, have been discussed by Ramsey (1956)and by Van Vleck (1957). Interestingly, negative tem-peratures are hotter than positive ones. Later, nuclearcooperative phenomena at T.0 and at T,0 were in-vestigated extensively by Abragam and Goldman andco-workers (Abragam and Goldman, 1982) in dielectricmaterials like CaF 2 and LiH and by Wenckebach andco-workers (Van der Zon, Van Velzen, and Wenck-ebach, 1990) in Ca(OH) 2. These studies, however, werelimited to ordering by the truncated dipolar force be-cause adiabatic demagnetization took place in a rotatingcoordinate frame. The main weakness of the method(see Sec. XIV) is, however, the inevitable presence ofelectronic paramagnetic impurities, introduced pur-posely for dynamic nuclear polarization by the ‘‘solideffect.’’ The local fields produced by the impuritiesprobably blur, to a certain extent, some of the featuresof the nuclear long-range order (Abragam, 1987).

Copper, silver, and rhodium, cooled in the laboratoryframe by the more general ‘‘brute force’’ method (Lou-nasmaa, 1974), without recourse to electronic impurities,have provided the most general and interesting systems,so far, for studies of nuclear magnetism near the abso-lute zero, at T.0 and T,0 (Hakonen, Lounasmaa, andOja, 1991). The Helsinki group has produced negativespin temperatures in silver and in rhodium.

Negative temperatures are more difficult to achieve inmetals than in insulators for two reasons: first, substan-tial effort is needed to reach the high initial spin polar-izations and, second, eddy currents make the productionof inverted spin populations difficult. First evidence forspontaneous nuclear magnetic ordering and for a nega-tive spin temperature was found by Oja, Annila, andTakano (1991) in silver. Better data were subsequentlyobtained by Hakonen and Yin (1991). Later, populationinversion from T.0 to T,0 was achieved very success-fully in silver and rhodium (Hakonen, Nummila, Vuo-

Page 60: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

60 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

rinen, and Lounasmaa, 1992; Hakonen, Vuorinen, andMartikainen, 1993). These feats were accomplished atultralow temperatures by reversing the magnetic fieldB quickly, in a time t!t2510 ms, so that the nuclei hadno chance to rearrange themselves adiabatically amongthe energy levels. In a certain sense, the transition frompositive to negative temperatures occurs viaT51`→2` , without crossing the absolute zero.Therefore the third law of thermodynamics is not vio-lated.

Rhodium nuclei have been refrigerated to 280 pK andto 2750 pK; these are, respectively, the current low- and‘‘high’’-temperature world records on positive and nega-tive sides of the absolute zero (see Sec. XVI.C). Produc-tion of negative temperatures has not succeeded in cop-per owing to the short spin-spin relaxation time, t25150 ms, of this metal.

At negative temperatures, nuclear spins provide newmodels for studies of magnetism. For example, in a fer-romagnet at T,0, the presence of long-range dipolarforces gives rise to the formation of magnetic domains ofa kind not found at T.0 (Abragam and Goldman,1982). The situation is more diverse in metals than ininsulators (Viertio and Oja, 1992). Since an externalmagnetic field does not provide a symmetry axis inbrute-force-cooled Cu, Ag, or Rh, in contrast to CaF 2and LiH cooled in the rotating frame, the spin structuresin these metals have a large degeneracy. A broad spec-trum of domain configurations are expected to occur.

It has sometimes been argued, that negative tempera-tures are fictitious quantities because they do not repre-sent true thermal equilibrium in a sample consisting ofnuclei, conduction electrons, and the lattice. However,the experiments on silver, in particular, show conclu-sively that this is not the case: the same interactions pro-duce ferro- or antiferromagnetic nuclear order in silver,depending on whether T,0 or T.0 (see Sec. VIII.C.3).Besides, true equilibrium, in the strictest sense of theword, hardly ever exists in nature. Furthermore, thelarge difference between the temperatures of conduc-tion electrons and of nuclei has no effect on the nuclear-spin structures.

B. Nuclear magnetic susceptibility of silver

The first measurements of the magnetic susceptibilityof silver down to the ordering temperature at T.0 weremade by Oja, Annila, and Takano (1991). The suscepti-bility was measured only as a function of spin entropy;the spin temperature was not determined. For this rea-son, and also because the susceptibility scale was notcalibrated, the data did not give much information aboutthe thermodynamics of the spin system.

These workers first realized and experimentally dem-onstrated that negative spin temperatures in silver canbe reached by a rapid field reversal (see Fig. 63). NMRemission, instead of absorption, showed that T,0. Thefield flip caused, however, a large loss of polarization.

The work on silver was continued subsequently byHakonen and Yin (1991). They performed extensive

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

measurements of the magnetic susceptibility and en-tropy of silver nuclei down to the extremely low spintemperature of 0.8 nK and, at T,0, up to 24.3 nK.These temperatures are much lower than those reachedin copper (see Sec. V.A) because the interactions areweaker in silver. The highest initial polarizations in theexperiments were p50.72 at T.0 and p5 –0.40 atT,0. The work has been described in considerable de-tail by Hakonen and Yin (1991).

All magnetic-susceptibility measurements were madeon a bundle of polycrystalline silver foils. A typicalspecimen weighed 2 g and consisted of 78 foils, 25 mm34.5 mm 340 mm along the x , y , and z directions, re-spectively. The material, of 99.99% nominal purity, wasselectively oxidized at 750 °C for 20 h in 0.1 mbar pres-sure of dry air to neutralize the magnetic impurities,which otherwise might have shortened the spin-latticerelaxation time t1 in small magnetic fields. The heattreatment and oxidization increased the residual resistiv-ity ratio (RRR) of the silver sample from 100 to 900.Details of sample preparation were discussed in Sec.III.D.

The dynamic susceptibility x8(f)2ix9(f) of the silverspecimens was measured by using low-frequencySQUID-NMR techniques, which were described in Sec.III.E. Two coaxial mu-metal cylinders were used asmagnetic shields around the silver foils to exclude the63-mT remanent field of the main demagnetization so-lenoid. Inside the shields, three small coils were as-

FIG. 63. Production of a negative spin temperature in silver.(a) Initial NMR spectrum at 0.77 mT. (b) Spectrum at –0.77mT after a quick reversal of the field. The negative absorption,i.e., emission of the 109Ag spins indicates that T109,0. How-ever, T107 still appears to be positive, perhaps owing to a verylow temperature Tss of the interaction reservoir (see Fig. 6).(c) Spectrum after lowering the field to zero and bringing itback to –0.77 mT. Now T1075Tss5T109,0. From Oja, Annila,and Takano (1991).

Page 61: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

61A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

sembled for NMR experiments, as illustrated in Fig. 16.Most NMR measurements were performed in a steadyfield B5Byy along the y direction, i.e., parallel to the4.5-mm-long edge of the sample foils. This field was pro-duced by a saddle-shaped coil while the rf excitationfield Bz

rf was generated by a solenoid. An astatic pair ofcoils was used for pickup (see Fig. 16).

The spin temperature was measured as described inSec. IV. Polarization was determined from the NMR ab-sorption x9, recorded in a field B@B loc535 mT, by us-ing the equation p5A*x9(f)df , where f is the NMRexcitation frequency. A small field, By5191 mT, waschosen for the polarization measurements because theseparate isotopic identity of the spins is then largely lost,which results in a single exchange-narrowed NMR line,integrable with good precision. The accuracy of the po-larization calibration was estimated as 5%, at both posi-tive and negative spin temperatures.

NMR data on silver at T51 nK and – 4.3 nK (Ha-konen, Yin, and Lounasmaa, 1990) are displayed in Fig.64; x9 was measured by increasing the frequency at therate df/dt51 Hz/s. At T,0, x9 was negative, indicatingthat, instead of absorbing, the spin system was emittingenergy. The magnitude of x9 was almost three timeshigher at negative temperatures because of the large fer-romagnetic susceptibility. The peak frequency of the

FIG. 64. NMR absorption and emission spectra for silver mea-sured in zero magnetic field: s , at T51.0 nK; d , at T5 –4.3 nK. Solid curves are fits of Lorentzian line shapesxL9(f)2xL9(2f)5A/@11(f2f0)2/G2#2A/@11(f1f0)2/G2#to the experimental data, with A50.457, f0539.6 Hz, andG559.6 Hz at T51 nK and with A5 –0.911, f0519.8 Hz,and G543.8 Hz at T5 –4.3 nK, respectively. Note the differ-ent vertical scales for T.0 and T,0. G is the half-width of theNMR peak. From Hakonen, Yin, and Lounasmaa (1990).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

NMR spectra is clearly lower at T,0 than at T.0. Theexperimental points fit nicely to Lorentzian lineshapes.

The nuclear entropy, calculated from polarization [seeEq. (26a)], is shown in Fig. 65 as a function of uTu. Thehigh-temperature expansion fits the data above 5 nK,but clear deviations from the 1/T2 law are observed be-low. Within the experimental accuracy, the results atT.0 and at T,0 coincide.

In Fig. 66, the absolute value of the inverse staticsusceptibility u1/x8(0)u of silver, calculated from theKramers-Kronig relation [see Eq. (27)] x8(0)5(2/p)*(x9/f)df by integrating from 30 to 180 Hz, isplotted as a function of uTu in nanokelvins. Clearly, thesusceptibility at T,0 is much larger than at T.0. Thisresults from the fact that the spin system tries to maxi-mize its energy at constant entropy when T,0, as dis-cussed in Sec. II.D.3. Since the exchange interaction isantiferromagnetic in silver, the maximum-energy state

FIG. 65. Reduced entropy 12S/R ln2 vs the absolute value ofthe spin temperature in silver: s , T.0; d , T,0. The dashedline displays the leading 1/T2 term of the high-T expansion.From Hakonen, Yin, and Lounasmaa (1990).

FIG. 66. Absolute value of the inverse static susceptibilityu1/x8(0)u vs the absolute value of temperature for silver: s ,measured at T.0; d , measured at T,0. Straight lines corre-spond, respectively, to the antiferromagnetic and ferromag-netic Curie-Weiss laws. The external magnetic field B50.From Hakonen, Yin, and Lounasmaa (1990).

Page 62: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

62 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

has a ferromagnetic alignment of spins, and the suscep-tibility is large.

The data in Fig. 66 display an antiferromagneticCurie-Weiss law, x5CAg /(T2uA) at positive tem-peratures, with uA5 –4.8 nK; CAg52 nK is the Curieconstant. At T,0, a ferromagnetic law, uxu5CAg /(uTu2uF) with uF52.8 nK, was obtained. Ac-cording to the mean-field theory, uuAu5uuFu (see pp. 715– 718 in Ashcroft and Mermin, 1976). By averaging thetwo measured values and taking into account the Lor-entz and demagnetizing factors [see Eq. (34)], one findsu/CAg5 –2.5 in accordance with data obtained usingspin dynamics (Oja, Annila, and Takano, 1990).

It is worth noting that both sets of results in Fig. 66follow the Curie-Weiss law to the lowest temperatures.The susceptibility displays no saturation as would be ex-pected close to an antiferromagnetic transition tempera-ture when T.0 (Huiku et al., 1986). Furthermore, in themeasurements by Hakonen, Yin, and Lounasmaa(1990), no changes were observed in the NMR line-shapes that could be assigned to actual antiferromag-netic or ferromagnetic ordering at T.0 or at T,0, re-spectively. Clearly, the transition temperatures were notreached in these early experiments.

C. Nuclear ordering of silver at T>0

1. Search for the transition

The first signs of spontaneous nuclear ordering in sil-ver were seen in the experiments of Oja, Annila, andTakano (1991). Spin entropies down to 0.50R ln2, cor-responding to polarizations up to p50.78, were ob-tained. Indications of antiferromagnetic ordering weredetected in the static magnetic susceptibility, whichshowed clear saturation at the beginning of the warmupafter cooling to T,1 nK (see Fig. 67). The amount ofuseful data gathered in these experiments was, however,somewhat marginal.

More extensive and successful measurements then fol-lowed (Hakonen and Yin, 1991; Hakonen, Yin, and

FIG. 67. Static susceptibility x8(0) of silver as a function ofpolarization p , which decreases with time after demagnetiza-tion. The different symbols refer to three separate cooldowns.From Oja, Annila, and Takano (1991).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Nummila, 1991). The experimental procedure and thepreparation of the specimens were the same as alreadydescribed, except for a simple but important improve-ment: folding of the sample foils (see insert in Fig. 15) toincrease rigidity, which resulted in a decrease of vibra-tional heat produced in the specimen. The highest polar-ization reached by the silver nuclei was p50.94, whichcorresponds to a 190 mK nuclear-spin temperature inthe 7.3-T field.

The clearest indication of antiferromagnetism in silvercame from changes in the NMR spectra upon ordering.The results depended on the way in which the demagne-tization was performed. Usually the magnetic field wasreduced directly to its final value. Figure 68 depicts fourexamples of the recorded line shapes; the dashed curvecorresponds to the paramagnetic state. The spectra dis-play an upward frequency shift after demagnetization,especially in low fields.

The most important observable in these experimentswas, however, again the static susceptibility,x8(0)5(2/p)*(x9/f)df , obtained by integrating overthe measured NMR lines. Figure 69 illustrates the evo-lution of x8(0) with time. The data, which correspond toa steady increase of temperature and entropy, displayseveral features that may be associated with a magnetictransition. The fast initial decrease of susceptibility at 2mT was probably caused by the rapid disappearance ofthe supercooled paramagnetic phase. Next, in smallfields, a slight increase of x8(0) was observed before amaximum, marked by an arrow, was reached; a mono-tonic decrease of susceptibility then started. In higherfields, there was only a kink in the x8(0) vs t curve, andsometimes no transition at all could be identified.

In analogy with electronic magnets, the maximum ofsusceptibility in small fields was regarded as an indica-tion of antiferromagnetic order in the nuclear-spin sys-tem of silver. The kink was identified approximatelywith the Neel temperature TN . The values pN50.76560.02 and TN5(560660) pK were deducedfrom these experiments. Since demagnetization of the

FIG. 68. NMR spectra of silver nuclei measured 5 min afterdemagnetization had been completed to the final field denotedin each frame. Dashed lines depict data recorded in the para-magnetic phase about 1 h later. For further explanations, seetext. From Hakonen and Yin (1991).

Page 63: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

63A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

sample was adiabatic, critical polarization could be con-verted to critical entropy using the equations of theparamagnetic state [see Eq. (26a)]; the resulting phasediagram will be discussed in Sec. VIII.C.3.

2. Magnetic preparation of the sampleby the demagnetization scheme

Hakonen and Yin (1991) also conducted experimentsin which the sample was ‘‘prepared’’ by first demagne-tizing it to a small reversed field, –5 mT, before thesweep to the final field. These measurements were ini-tially motivated by the strange shape of the phase dia-gram, to be discussed in Sec. VIII.C.3, which suggestedthat the ‘‘real’’ ground state was hidden by a large hys-teresis with respect to the external field. Some of theresults are illustrated in Fig. 70. In small final fields, theabsorption at low frequencies was enhanced as shown bythe first curve. When the final field was 24 mT, the NMRabsorption split into two resonances, one below and theother slightly above the peak in the paramagnetic state.There was a transfer of intensity from the low-frequencypeak to the high-frequency mode as the spins warmedup. The line shapes were fairly reproducible, althoughthe actual spectral weights of the peaks varied from runto run.

The frequency of the low-f mode, fd , obtained frommeasurements in which the sample was prepared in areversed field, is independent of B , as is illustrated byFig. 71. The position of the upper absorption peak, fu , isalso shown, as well as the location of the paramagnetic

FIG. 69. Static susceptibility of silver nuclei as a function oftime after demagnetization to four different external magneticfields. Each set of data is scaled by the maximum susceptibilityxmax8 (0) for that run. Small arrows indicate the transition pointfrom the ordered to the paramagnetic phase. From Hakonenand Yin (1991).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

FIG. 70. NMR absorption spectra of silver, after the sample-preparation sequence B5190 mT→–5 mT→24 mT, measuredevery 15 min. The topmost curve, however, was recorded atB5 –5 mT, just before the spectrum below. The lowest curvecorresponds to the paramagnetic state. The successive spectraare shifted downwards along the vertical axis as shown. FromHakonen and Yin (1991).

FIG. 71. Frequencies of the NMR peak amplitudes fp vs theexternal magnetic field for silver. The antiferromagneticbranches are d , fu and s , fd . The peak position in the para-magnetic region (n) was determined slightly above TN .Dashed line, Larmor frequency f5gB/2p , averaged over thetwo silver isotopes. The experimental data were taken fromHakonen, Yin, and Nummila (1991). Solid lines, calculated po-sitions of antiferromagnetic resonances in a single-k structureas given by Eq. (153) (Heinila and Oja, 1996).

Page 64: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

64 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

peak. The upper absorption peak fu was determinedfrom spectra recorded when demagnetization proceededdirectly to the final field. The position of the upper peakin experiments with magnetic sample preparation wasslightly higher in frequency.

When the line shapes illustrated in Fig. 70 are inte-grated using the Kramers-Kronig relation [see Eq. (27)],one obtains the susceptibility vs time dependency shownin Fig. 72. It can be seen that x8(0) decays continuouslywith time; the behavior is very different from the casewhen the final field B524 mT was reached directly (seeFig. 69). Furthermore, the initial susceptibility is about20% higher than in measurements during which demag-netization proceeded directly to the final field. Signs of aplateau in the x8(0) vs t curves could be seen, however,in experiments with magnetic preparation to final fieldslower than 10 mT or higher than 50 mT.

Another feature that is difficult to understand is that,based on measurements illustrated in Fig. 69, the lowestNMR absorption curve in Fig. 70 corresponds to theparamagnetic state, although the enhanced absorption atlow frequencies suggests that antiferromagnetism is stillpresent.

Hakonen and Yin (1991) discussed various possibili-ties to explain these findings. The most natural reasonfor the decay of x8(0) with time would be the coexist-ence of a supercooled paramagnetic phase together withan ordered state, since the susceptibility of the paramag-netic phase well below TN is higher than the susceptibil-ity of the ordered state. Another possibility would bethat the spin structure obtained after sample preparationin a reversed field is a nonequilibrium ordered state, or anonequilibrium combination of different domains, witha slow relaxation towards equilibrium. The enhancedlow-f part of the NMR spectrum above TN is a problem,however. It is difficult to understand why the spectrumshould deviate from the paramagnetic shape.

3. Antiferromagnetic phase boundary

Figure 73 (see also Fig. 4) summarizes the NMR re-sults on magnetic ordering in silver and displays thephase diagrams for the ferro- and antiferromagneticstates in the B-S plane (Hakonen, Nummila, Vuorinen,

FIG. 72. Static susceptibility of silver vs time obtained fromthe NMR spectra shown in Fig. 70. From Hakonen and Yin(1991).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

and Lounasmaa, 1992). At T.0, the phase boundarybetween the paramagnetic and antiferromagnetic phasesis of second order in high fields. In small fields, belowapproximately 30 mT, the transition is presumably offirst order, as is suggested by the fact that the positionsof the antiferromagnetic features in the NMR lineshapesdo not change when the spins warm towards TN .

The shape of the phase diagram at T.0 in Fig. 73 isinteresting: silver seems to order magnetically more eas-ily in a small field than in B50. The matter has beendiscussed at some length by Hakonen and Yin (1991).If the phase diagram in the B-T plane is similar in formto that in the B-S plane, as in Fig. 4, the Clausius-Clapeyron equation of magnetic systems, dB/dT52(Sp2So)/(Mp2Mo), would imply that Mo , themagnetization in the ordered state, is larger than Mp inthe paramagnetic phase, since thermodynamic stabilityrequires that Sp.So . This would suggest ferromag-netism. All susceptibility data, however, point towardsantiferromagnetism at T.0. Therefore the phase dia-gram in the B-T plane must have the conventionalshape with dB/dT,0. This means that the isentropesmust be nonmonotonic, which agrees with the crossingof the measured entropies in the paramagnetic phase asillustrated in Fig. 74.

Hakonen and Yin (1991) directly measured TN inB532 mT. They found pN50.6760.02 and TN5(700660) pK. Since TN5(560660) pK in B50, the shapeof the phase boundary in the B-T plane should be simi-lar to that in the B-S plane. However, the measuredNeel temperatures in B50 and 32 mT were not consid-ered convincing proof of a positive dB/dT in low fields

FIG. 73. Phase diagram of nuclear spins in silver at T.0 (fullcurve) and at T,0 (dashed curve) in a magnetic field (B) vsreduced entropy (S/R ln2) plane. Inside the solid curve, thespin system is antiferromagnetically ordered. The experimen-tal points at T.0 (s) are from the paper by Hakonen, Yin,and Nummila (1991). The dashed curve for T,0 displays thephase boundary between the ferromagnetic domain state (in-side) and the paramagnetic region. The curve was determinedby the two data points and its intercept with the S50 axis. Theshape of the dashed curve is based on the mean-field theory[see Eq. (93)] by assuming a linear relationship between S andT . Modified from Hakonen, Nummila, Vuorinen, and Lounas-maa (1992).

Page 65: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

65A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

since, in addition to the error margins of TN quotedabove, there was the possibility of an additional 10%systematic error owing to the relative calibrations of pbetween 0 and 32 mT.

In their discussion, Hakonen and Yin (1991) mentionthe additional possibility that silver nuclei could displaysome exotic non-Neel-type short-range order, similar tothat discussed by Anderson (1973). If this were the case,this suggestion would provide a new interpretation forseveral observations, like the shape of the phase dia-gram, the decay of x8(0) with time (see Fig. 72), and theNMR line shapes of Fig. 70.

4. Number of antiferromagnetic phases

It is difficult to decide, on the basis of experimentaldata, how many antiferromagnetic structures there arein the magnetically ordered region of silver spins. Theexplanation preferred by Hakonen, Yin, and Nummila(1991) is that there is only one antiferromagnetic phasein the whole ordered region, a single-k structure. Thetwo antiferromagnetic peaks, centered at frequenciesfu and fd (see Fig. 71), would then correspond to the twoantiferromagnetic resonance modes of a single-k struc-ture with type-I order (Kumar et al., 1986; Heinila andOja, 1996). The enhancement of the fd mode after anexcursion to a small reversed field could result from areorientation of spins in the various ordered crystallites.

In principle, the fu and fd modes may also result fromtwo different ordered phases, for example, single-kstructures, and the intensity of the fu and fd modes couldreflect the relative proportions of the two coexistingphases.

Hakonen, Yin, and Nummila (1991) also point outthat the behavior of the antiferromagnetic absorptionpeaks can be used to argue for a phase transition infields on the order of B550 mT. As is shown by Fig. 71,the shift of the upper antiferromagnetic mode at f5fu

FIG. 74. Reduced entropy 12S/R ln2 of silver vs tempera-ture: s , B50; d , B532 mT. Note the crossing of the twocurves; the inset illustrates the shape of the isentropes S1 andS2 (.S1). The dashed lines display the leading 1/T2 terms ofthe high-temperature expansions. From Hakonen and Yin(1991).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

from the paramagnetic line vanishes around this field,and the intensity of the lower mode becomes very small.This transition would have to be of second order.

The NMR behavior of silver is distinctly differentfrom that of copper. In polycrystalline Cu, Huiku et al.(1986) found two phases that displayed different reso-nance frequencies (see Fig. 68): a low-field peak belowthe paramagnetic resonance and a high-field peak abovethe paramagnetic line. In silver, similar frequency shiftswere observed but in small fields only. In addition, thefrequency of the fu mode in silver depends on B , incontrast to the behavior observed in copper.

5. Comparison with theory

At T.0, the general features of ordering seem toagree with theoretical predictions. The measured transi-tion temperature, 560 pK in zero field, agrees well withthe value, 500 pK, obtained from Monte Carlo simula-tions by Viertio (1990). The first-order nature of thetransition in small fields is reproduced by these calcula-tions as well, but the critical entropy was not computed.The theoretically calculated critical field, Bc5140 mT, issomewhat higher than the extrapolated experimentalvalue of 100 mT. The prediction for TN depends sensi-tively, however, on the spin I . The Monte Carlo resultswere obtained for classical spins. An estimate for thequantum correction due to the spin I5 1

2 of silver nuclei,to be discussed in Sec. XV.C.3.c, yields a factor-of-3 toohigh ordering temperature, TN51.5 nK. The same TN isalso predicted by the spherical model calculation of Har-mon et al. (1992).

The spin-spin interactions of silver have been com-puted from the first-principles electronic band structure(Harmon et al., 1992). These calculations will be dis-cussed in Sec. XV.A. The overall strength of the theo-retically calculated spin-spin interaction is in goodagreement with the values obtained from susceptibilitydata (Hakonen, Yin, and Lounasmaa, 1990) and fromNMR measurements (Poitrenaud and Winter, 1964; Oja,Annila, and Takano, 1990; Hakonen, Nummila, andVuorinen, 1992).

Comparison of the measured susceptibility againstspherical-model calculations gives the best fit if thestrength of the RK interaction is decreased in absolutevalue from R522.3 to R521.7 (Harmon et al., 1992).A weaker coupling would also decrease the discrepancybetween the experimental and theoretical ordering tem-peratures.

The magnetic phase diagram of silver has been inves-tigated in several calculations. One might naively expectthat there is only a single ordered state because theforces between silver atoms are rather isotropic sincethe Ruderman-Kittel interaction dominates. It turns out,however, that for a type-I fcc antiferromagnet with onlyisotropic, nearest-neighbor interactions, there is a tran-sition between a low-field single-k structure and a high-field triple-k state (Heinila and Oja, 1993a, 1994b; seeSec. XV.D.6). For isotropic interactions, the spin struc-tures do not depend on the direction of the field with

Page 66: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

66 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

FIG. 75. NMR emission x9 and dispersionx8 curves vs frequency for the nuclear-spinsystem of silver at T,0, measured in B50for reduced entropies: s , S/R ln250.87; n ,S/R ln250.80; h, S/R ln250.73. These en-tropies correspond in the paramagnetic stateto polarizations p 520.42, 20.51, and20.59, respectively. In both frames, the zerosfor the two lower curves have been shifteddownwards by 0.4 and 0.9 units, respectively.From Hakonen, Nummila, Vuorinen, andLounasmaa (1992).

respect to crystalline axes, which is the case when thedipolar interaction between silver nuclei, albeit rela-tively small, is taken into account. The ground-state cal-culations including dipolar interactions are discussed indetail in Sec. XV.D.7. The overall conclusion from vari-ous theoretical models is that there should be a single-k→ triple-k transition in the field interval B/Bc50.3 –0.6 when B is aligned along the [001] or [110]crystalline axis, but that when B is parallel to [111] atriple-k structure exists only in a small region at inter-mediate fields (Viertio, and Oja, 1987; Viertio, 1992;Heinila and Oja, 1993a). In addition, a transition be-tween two different single-k structures is possible in lowfields. However, neutron-diffraction measurements ofTuoriniemi, Nummila et al. (1995) on silver (see Sec. IX)revealed only one single-k structure in fields below Bcwhen Bi[001].

The NMR frequencies of type-I fcc antiferromagnetshave been calculated by Kumar et al. (1986) by solvingthe mean-field equations of motion. By comparing thepeak positions fu and fd of the observed antiferromag-netic resonances with theoretical predictions for asingle-k structure, Hakonen, Yin, and Nummila (1991)concluded that such a configuration can describe thedata in the whole ordered region. The matter has beenstudied theoretically in more detail by Heinila and Oja(1996), who also conclude that the observed antiferro-magnetic resonances can be accounted for by a single-kstructure. For more details, see Sec. XV.E.2.

D. Nuclear ordering in silver at T<0

1. Observation of the ferromagnetic transition

The final success, nuclear ordering in silver at T,0,came in 1991 (Hakonen, Nummila, Vuorinen, and Lou-nasmaa, 1992). In order to facilitate the production of

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

negative temperatures by rapid field reversal, a new coilarrangement and improved radiation shields were care-fully prepared to prevent eddy currents, which musthave been the main problem during previous, less suc-cessful experiments. Two saddle-shaped static field coilsBx and By and a solenoid Bz were assembled on a coilformer inside the brass radiation shield, which had fourcuts along its length (see Fig. 53). An astatically woundpickup coil was oriented parallel to the z axis.

The final phase of the demagnetization sequence (seeSec. III.C) was modified, as well. Degaussing of thesecond-stage magnet was employed: +7.4 T→ –0.1 T→+0.02 T→ 0. This guaranteed proper operation of theSQUID measurement system, which was susceptible tovibrational noise in the presence of even a small rema-nent field between 0 and 5 mT. Fortunately, the field-cycling procedure did not substantially decrease thelargest initial polarization, p50.85, achieved in theseexperiments. The external demagnetization field was inthe vertical direction, B5Bz.

The solenoidal magnetic field Bz , employed in therapid field reversal, was typically 400 mT. The best in-version efficiency, 90%, was reached for small initial po-larizations on the order of p50.3. At higher polariza-tions, the results varied irregularly. Despite many effortsby changing the speed, symmetry, and magnitude of thefield reversal, the optimum inversion efficiency was only60–75% at p50.85. Polarizations at T,0 were thuslimited to p'20.65.

Figure 75 illustrates NMR line shapes measured inzero field at T,0. Instead of absorption, the spin systemis again emitting energy (see Fig. 64). The emissionmaximum shifts towards higher frequencies with in-creasing upu. The solid curves are simultaneous fits of theemission and dispersion curves to Lorentzian lineshapes.

Page 67: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

67A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Figure 76 displays the perpendicular static susceptibil-ity x'5x8(0), integrated from the Kramers-Kronig re-lation [see Eq. (27)], as a function of polarization in zerofield and at four different values of By . In addition, dataobtained in the parallel field Bz55 mT are also shown.In zero field and at 5 mT, there is first a monotonic in-crease of x8(0) with p , which then saturates in the re-gion p520.4 – 20.5. This behavior is caused by ferro-magnetic ordering; the susceptibility is governed bydipolar interactions through the formation of ferromag-netic domains, which, in the ideal case, would lead tox'521 in the silver sample. Two possible configura-tions are shown in Fig. 77.

In zero field and at small polarizations the suscepti-bility of silver is of the form x8(0)5x0 /

FIG. 76. Static susceptibility x8(0) vs polarization p of silvernuclear spins at T,0: s , measured in a magnetic field By

50, ,, By55; h , By520; L , By550; n, By5100 mT; 3 ,Bz55 mT. The scale on top gives the estimated temperaturebased on the formula 1/upu2150.55 uT/nKu, which applies atB50; values to the right of 22 nK are only suggestive. FromHakonen, Nummila, Vuorinen, and Lounasmaa (1992).

FIG. 77. Two examples of proposed (Viertio and Oja, 1992)domain configurations in flat silver specimens. The single-domain structure at left is possible only when M is perpendicu-lar to the surface of the foils. In the multidomain structure atright, M is parallel to the foils.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

@12(R1L2Dz)x0# , [see Eq. (34)] where x051.30p isthe Curie-law value for a noninteracting spin system andthe Lorentz factor L5 1

3. The uppermost curve in Fig. 76is the best fit of points in the region upu,0.45 by thisequation, using R521.1, which differs significantly fromthe value R522.560.5 obtained from other NMR ex-periments (see Sec. XI.B.2). This indicates that substan-tial deviations from the mean-field behavior are alreadytaking place at intermediate polarizations. Within thescatter of the measured data, the same curve fits theexperimental results at B50 and at B55 mT. As is typi-cal for dipolar ferromagnetism (Abragam and Goldman,1982), the data do not saturate completely even whenthe polarization is high but, nevertheless, it was possibleto describe the results for upu.0.55 approximately by aconstant value, x5xsat521.05.

Hysteresis is often a characteristic feature of ferro-magnetic ordering, best seen in a x8 vs B plot. Figure 78depicts transverse ac susceptibility at f510 Hz (Ha-konen and Vuorinen, 1992) when the field was sweptback and forth between 630 mT at p520.65. The dip inthe middle is due to the fact that the ratio x8(10 Hz)/x8(0) increases with the field. On the basis of Fig. 78and similar plots, no hysteresis was found in the ferro-magnetic phase of silver. This is consistent with thetheoretical structure discussed earlier: Because of thelarge degeneracy in the spin directions, the magnetic do-mains can adjust smoothly to variations in the externalfield (Viertio and Oja, 1992).

2. Phase boundary of the domain state

Figure 73 (see also Fig. 4) summarizes the results onmagnetic ordering in silver and displays the phasediagrams for the ferro- and antiferromagnetic states in

FIG. 78. Transverse ac susceptibility 2x8(10 Hz) of silver,measured at p5 –0.65 while sweeping the magnetic field backand forth: s , sweeps to the left; d , sweeps to the right. B wasoriented perpendicular to the sample foils. From Hakonen andVuorinen (1992).

Page 68: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

68 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

the B-S plane (Hakonen, Nummila, Vuorinen, and Lou-nasmaa, 1992). At T,0, there are experimental pointsonly at zero and 5-mT fields; at 10 mT the transitioncould not be identified reliably from the susceptibilityexperiments (see Fig. 76) on which Fig. 73 is based. Thedashed line sketches the stability region for the domainstates when T,0. The curve was obtained by applyingthe mean-field theory, together with the estimateBc(T50)52m0Msat /xsat'40 mT, calculated by usingthe observed saturation value of the susceptibilityxsat521.05.

The crossing of the two uppermost lines in Fig. 76 wasidentified as the transition point to the ferromagneticstate. Owing to the rounding of the x8(0) vs p curve, aconservative estimate for the critical polarization, inzero field and at B55 mT is pc5 2 0.4960.05, whichcorresponds to Sc /R ln250.8260.035. By employing thelinear relationship between the temperature and inversepolarization (see Fig. 21), 1/upu2150.55uT/nKu, one ob-tains for the Curie point TC5(21.960.4) nK.

3. Comparison with theory

At negative spin temperatures, according to themean-field theory, TC5u5(R1L2D)C525.7 nK insilver, which is clearly different from the observed value;here we have used R522.2, L5 1

3, and D51.10 Calcu-lations on the spin-1

2 Heisenberg model (De Jongh andMiedema, 1974) in an fcc lattice yield TC50.67u50.67 1

3I(I11)12J . Use of the theoretical (see Table V)nearest-neighbor interaction J/kB521.3 nK givesTC522.6 nK, while u525.7 nK yields 23.8 nK. Sincethe latter result takes into account the dipolar interac-tion as well, it is clear that the experimental TC is sig-nificantly lower than theoretical estimates. In addition,the critical entropy in the Heisenberg model,Sc50.66R ln2, is clearly lower than the experimentalvalue 0.82R ln2. This is reasonable since the long-rangedipolar force, which was not included in the calculation,acts ferromagnetically at negative temperatures andshould increase Sc .

Monte Carlo calculations by Viertio and Oja (1992)predict TC5 –1.7 nK, which is close to the measuredresult. These computations were made, however, assum-ing classical spins. An estimate for the correction due tothe spin I5 1

2 of silver nuclei multiplies the predictedordering temperature by a factor of 3, thus yielding TC5 –5.1 nK (see Sec. XV.C.3.c). This theoretical value isin approximate agreement with the result obtained bycorrecting the Curie temperature of De Jongh andMiedema (1974) with the dipolar interaction. MonteCarlo simulations yielded Sc50.93R ln2, which deviatessubstantially from the experimental value. This is notsurprising, since a too high Sc is likely to ensue when aquantum-spin assembly is modeled using a classical sys-tem. It seems, therefore, that there remains a serious

10TC should be computed for D50 at T.0 and for D51 atT,0.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

discrepancy between the measured and predicted order-ing temperatures in silver, although there is at least arough agreement for critical entropies.

The saturation of susceptibility to xsat'21 in the or-dered state at T,0 can be explained only by the forma-tion of domains, since otherwise the susceptibility woulddiverge at TC . The domain configurations are such thatthey maximize the magnetic enthalpy H5U2BM . In-stead of needles, as at T.0, platelike domains are ex-pected when T,0 (Abragam and Goldman, 1982).There is a rich variety of different energetically degen-erate domain configurations (Viertio and Oja, 1992).These will be described in detail in Sec. XV.H. Acommon feature of the different domain structures isthat the susceptibility along the external field isxsat5m0M/B521/(12DM), where DM is the demagne-tization factor along B. This is in agreement with theexperimental data at B50 and 5 mT. The observed be-havior is clearly different from ferromagnetic orderingat positive temperatures; in this case the susceptibilitysaturates at xsat51/D . The difference is caused by theshape of ferromagnetic domains in silver. At T,0, thetangential component of the magnetization is alwayscontinuous and the normal component changes sign at adomain boundary; this is illustrated in Fig. 4. The oppo-site is true at positive temperatures.

IX. NEUTRON-DIFFRACTION EXPERIMENTS ON SILVER

A. Experimental arrangements

By means of NMR measurements it is not possible toverify the details of the spin structure in silver. Experi-ments employing scattering of neutrons are necessaryfor this purpose. Tuoriniemi, Nummila et al. (1995) haverecently observed long-range nuclear antiferromagneticorder by neutron diffraction in a single crystal of silver.Further details of this research have been described byTuoriniemi, Lefmann (1995), Lefmann et al. (1995),Tuoriniemi (1995), Tuoriniemi et al. (1996), Lefmannet al. (1996), and Nummila et al. (1997). For this workthe sample had to be isotopically pure, since 107Ag(51.8%) and 109Ag (48.2%) in natural silver have oppo-site signs of the spin-dependent scattering lengths,strongly depressing the coherent neutron signal that in-dicates the alignment of nuclear spins. The experimentused a 99.7% enriched material of 109Ag to grow the 0.7312 325 mm3 single crystal. The [11 0] axis was parallelto the longest edge of the crystal, which was mountedupright in the cryostat. The cubic direction [001] pointed24° away from the flat surface of the sample, see Fig.79(a).

The experiments were performed at the BER II reac-tor of the Hahn-Meitner-Institut in Berlin, in the groupof Prof. M. Steiner. The diffracted neutrons were re-corded at a fixed scattering angle by a single counter orby an area-sensitive detector. Another neutron countermonitored the transmitted beam. The l/2 contaminationwas removed from the monochromatic beam by a beryl-lium filter.

Page 69: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

69A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

The sample was cooled in a cascade nuclear-demagnetization cryostat (see Sec. VI.B), which had a9-T magnet surrounding the massive copper coolingstage and a 7-T magnet for the sample. The copper re-frigerant, demagnetized to 60 mT, kept the lattice tem-perature at 100 mK while the 109Ag nuclei were polar-ized to about 95%. The silver nuclei were cooled furtherinto the picokelvin range by reducing the 7-T externalfield to zero. Before the end of demagnetization, an ad-ditional 500-mT field was applied to the sample by a setof small coils, so that the ordered state could be enteredfrom any field direction. During all stages of the experi-ment, the lattice and the conduction electrons of the109Ag sample remained in thermal contact with the cop-per nuclear stage. During measurements the neutronbeam was the main source of heat, reducing t1 from 12 hto 3 h. Prior to demagnetizations, the diffractometer wasaligned to the (0 0 1) Bragg position of a type-I antifer-romagnet in an fcc lattice.

B. Results

In Fig. 80 two sets of neutron-diffraction data inB50 are shown. The nuclei were demagnetized into theordered state with the external field B along the [001] or[010] directions, and neutron counts were monitoredwhile the spin system warmed up. A clear (0 0 1) reflec-tion appeared when demagnetization was made with Bparallel to the corresponding ordering vector. The pres-ence of this signal provided unambiguous proof of anti-ferromagnetic order in silver. It is important to note that

FIG. 79. Field-rotation observations of silver. (a) Orientationof the crystallographic axes within the slab-shaped silversample (shaded rectangle) is illustrated by a cube in the appro-priate alignment. (b) Field directions in which the (0 0 1) re-flection was observed during the field-rotation experiments areshown by heavy lines (the center of the cube is at the origin).The end points and their symmetry equivalents are marked byblack squares. A structure with k = (p/a)(0,0,1) was thus con-fined so that B had to be around [001] within a cone of 110°full opening; its outer surface is shaded in the figure. (c) Fielddirections, [001] and @0.8 0.81], in which the phase diagram wasstudied in greater detail; the crystal is viewed from above.From Tuoriniemi, Nummila et al., 1995.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

the neutron signal remained essentially absent when theordered state was entered from the perpendicular direc-tion [010], although in zero field the three ordering vec-tors, producing the (100), (010), and (001) reflections,are equivalent owing to the cubic symmetry. It was con-cluded that the observed antiferromagnetic state had asimple single-k structure and that the stable spin con-figuration was created during demagnetization.

To demonstrate that the observed intensity indeedwas a Bragg peak, Tuoriniemi, Nummila, et al. (1995)used an area-sensitive detector. Time development ofthe diffraction pattern, integrated over the vertical di-mension, is shown in Fig. 81. The line shape of the anti-ferromagnetic peak was Gaussian, and its width wascomparable to that of the (0 0 2) second-order latticereflection.

The critical entropy of ordering was found using theneutron-transmission data. The nuclear polarization pcould be deduced from the count rate when the nucleiwere aligned by a field in the paramagnetic state becausethe neutron absorption as well is spin dependent. Forthis purpose the 500-mT field was applied at the begin-ning of each experiment. The orientation of this fieldalso determined the direction along which the orderedstate was entered. Polarization was measured again afew times after the disappearance of the antiferromag-netic signal, and the critical value pc was found by inter-polation. The nuclear entropy S could then be calcu-lated from the paramagnetic polarization in 500 mT. Thefield changes were nearly adiabatic (DS;0.01R ln2 foreach sweep between B50 and 500 mT), whereby theentropy was known in any field. In the zero-field experi-ment of Fig. 80, pc50.75 6 0.02 was obtained, corre-sponding to Sc5(0.5460.03)R ln2.

FIG. 80. Time dependence of neutron intensity, measured by asingle counter (30 s per point) at the (0 0 1) position. The ini-tial polarization p50.9160.02 was recorded in a 500-mT fieldin the paramagnetic phase, after which B, in the [001] or [010]direction (d and s , respectively), was reduced to zero (65mT) at t53 min. The (0 0 1) neutron signal appeared immedi-ately below Bc , but only when Bi[001]. The inset shows theexponential relaxation of the nuclear polarization, with t153.1 h. The critical value pc50.7560.02 was found by inter-polation. From Tuoriniemi, Nummila et al., 1995.

Page 70: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

70 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

To estimate the Neel temperature of nuclear orderingin 109Ag, Tuoriniemi, Nummila et al. (1995) used thesemiempirical relation between T in zero field and p inthe paramagnetic state, 1/p21}T , established in earlierNMR measurements on natural silver (see Sec. IV.E):TN5(700680) pK. This is larger than 560 pK for natu-ral silver because the strength of the mutual interactionsis scaled by the magnetic moment squared, i.e., by afactor of 1.15.

The antiferromagnetic sublattice polarization pAF wasestimated by comparing the Bragg intensity I(001) withthe strength of the (0 0 2) second-order lattice peak. Themaximum signal in Fig. 80, at p50.90, corresponds topAF50.4860.10. Extrapolation of the almost linear de-pendence of I(001) on (1/pc21/p) suggests that pAF atp51 (T50) is still clearly below complete polarizationpAF51. This may reflect quantum fluctuations or frus-tration in the fcc lattice.

It was interesting to examine the response of the spinsystem to an applied magnetic field and to its alignment.With Bi[001], the antiferromagnetic intensity decreasedsmoothly when approaching the critical field Bc . Thespins thus lined up continuously towards the increasingfield, as in the spin-flop phase of a weakly anisotropicantiferromagnet. No field-induced phase transitionswithin the ordered state could be identified, which is incontrast to the single-k to triple-k transition theoreti-cally predicted by Viertio (1992) and by Heinila and Oja(1993a). Repetitive field cyclings across the phaseboundary to the paramagnetic state did not produce anyappreciable hysteresis nor excess nonadiabaticity; there-fore the ordering transition was presumably of secondorder.

FIG. 81. Time evolution of the antiferromagnetic Bragg peakin a 30-mT field. The 2u dependence of scattered neutrons isplotted as a function of deviation from the (0 0 1) position. Thebell-shaped curves are Gaussian fits to counts collected duringsix-minute intervals. Only every second spectrum is shown. Forclarity, the successive spectra are offset vertically by 5 cts/min.The residual peak, visible in the lower curves, is the (0 0 2)lattice reflection of second-order neutrons leaking through theBe filter (transmittance 0.3% at l/2). The initial and criticalpolarizations in this experiment were 0.89 6 0.02 and 0.69 60.02, respectively. From Tuoriniemi, Nummila et al., 1995.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

The effect of the field direction was studied by rotat-ing B with respect to the crystal axes. In a turn extend-ing from @ 1 1 0] to [110], the (0 0 1) reflection was visiblewhen the magnetic field (B550 mT) was aligned be-tween the axes @ 1 1 1] and [111]. Within this arc (;110°),the neutron intensity did not vary substantially. Anotherfield rotation was made in a perpendicular plane, start-

ing from @ 12

12 1# and passing over [01 1] towards

[11 0]. The neutron signal disappeared about 10° beyondthe [01 1] axis, confining the observed antiferromagneticphase roughly within a cone of 110° full opening. This isillustrated in Fig. 79(b).

Further experiments were made in two different fielddirections within this cone: Bi[001] and [email protected] 0.81].The latter is close to the edge, about 7° from the @ 1 1 1]axis. These directions were symmetric with respect tothe shape of the crystal, about 24° away from the flatsurface [see Fig. 79(c)], so that the same correction forthe demagnetization effect could be applied. The resultswere compiled into an entropy-versus-field phase dia-gram, shown in Fig. 82. The observed neutron intensitiesduring the respective experiments in the two directionswere comparable. On the other hand, the critical en-tropy was systematically higher near the edge of thecone than along the central [001] axis.

C. Comparisons with theory

The neutron-diffraction data of Tuoriniemi, Nummilaet al. (1995) can be compared with theoretical work.First of all, the observed type-I ordering vector had beenpredicted for silver on the basis of measured and calcu-lated interaction parameters (Oja and Kumar, 1987;Harmon, Wang, and Lindgard, 1992). The ground-statespin structure in Ag had been studied by perturbationanalysis (Heinila and Oja, 1993a) and by Monte Carlosimulations (Viertio, 1992). Both methods indicated that

FIG. 82. Magnetic field vs entropy phase diagram of silver;d , B was along the [001] direction; s , B was along the @0.80.8 1] direction; L , earlier NMR results (Hakonen, Yin, andNummila, 1991; Hakonen and Yin, 1991) on a polycrystallinesample, shown for comparison. The curves are just guides forthe eye. From Tuoriniemi, Nummila et al., 1995.

Page 71: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

71A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

in Bi[001] a single-k configuration is stable in lowmagnetic fields below B;0.5Bc . A structure with k5(p/a)(0,0,1) was expected, in perfect agreement withthe experimental observations. In higher fields, however,a triple-k configuration was predicted. According to thesimulations this structure is stable only if B is within anarrow cone (,5°) around the [100]-type axes. Themeasurements did not provide any evidence for thetriple-k state, although it was searched for in field-sweepand rotation experiments.

The simulations suggested further that only onek-vector, k = (p/a)(0,0,1), appears if B lies between the[001] and [111] axes, apart from the triple-k region closeto [001]. This is in agreement with the experimentaldata, except for the proposed triple-k phase. When B isturned beyond the [111] axis, the structure should trans-form first into another single-k configuration, giving riseto a (1 0 0) or a (0 1 0) reflection. Closer to [110], an-other triple-k state should exist when B.0.5Bc . Thesepredictions have not been tested experimentally, andtheir verification requires measurements in other fieldorientations.

X. STUDIES ON RHODIUM

A. Nuclear magnetic susceptibility

The natural choice for the Helsinki group to investi-gate, after copper and silver, would have been gold.There is some prospect of this metal’s becoming a super-conductor around Te5100 mK (Buchal et al., 1982). Theanticipated difficulties in obtaining a sufficiently puregold specimen, however, made rhodium the actualchoice. One reason for selecting Rh metal was the pos-sibility of investigating the effect of superconductivity onthe nuclear-spin system. In rhodium the critical param-eters are Bc=4.9 mT and Tc=325 mK (Buchal et al.,1983). Strong supercooling, even after active compensa-tion of the remanent field down to 0.2 mT, may havebeen the reason why the Helsinki group did not observesuperconductivity, but, as a result of their study,rhodium now provides another nuclear-spin system inwhich negative temperatures have been produced (Ha-konen, Vuorinen, and Martikainen, 1993). The spin-spinrelaxation time in Rh, t2=10.5 ms, is almost the same asin silver.

For a Heisenberg model in an fcc lattice, with an ap-preciable next-nearest neighbor interaction, it is possiblethat both the minimum (T.0) and the maximum(T,0) energy states are antiferromagnetic. Rhodiumappears to be the first such metal known. In addition tothe dipolar and Zeeman energies, the Hamiltonian con-tains contributions from the isotropic Ruderman-Kittelinteraction and from the anisotropic, so-called pseudo-dipolar interactions (Bloembergen and Rowland, 1955).The theory of indirect exchange interactions is discussedin Sec. XV.A. There is only one stable isotope, 103Rh,with spin I5 1

2. Quadrupolar contributions to the Hamil-tonian are thus absent. In rhodium, the strength of theexchange interactions, as compared to dipolar forces, is

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

intermediate between those in copper and silver. It canbe concluded, on the basis of NMR linewidth measure-ments, that the exchange interaction is nearly isotropic(Narath, Fromhold, and Jones, 1966). Pseudodipolar in-teractions are either small in comparison with the dipo-lar forces or, alternatively, they effectively change thesign of the dipolar interaction. For detailed discussionswe refer the reader to Vuorinen, Hakonen, Yao, andLounasmaa (1995).

The first measurements on rhodium were made byHakonen, Vuorinen, and Martikainen (1993) using twospecimens. Both samples were assembled from thin foils.The first specimen contained about 100 ppm of iron im-purities, which were responsible for a decrease of theKorringa constant k (see Sec. II.B) from 10 sK to 0.06sK in small magnetic fields. In the second specimen, withless than 15 ppm of electronic magnetic impurities, k50.2 sK in zero field. After heat treatment at 1330 °C inan oxygen atmosphere of 0.4 mbar for 16 h, the residualresistivity ratios (RRR) of the two specimens were 250and 530, respectively, and their effective iron contents14 and 6 ppm. Most of the results reported were ob-tained using the purer Rh sample.

The experimental arrangement and the cooling proce-dure were basically the same as employed in earliernuclear-refrigeration experiments in Helsinki (see Sec.III.B.). The Rh sample again formed the second stage ofthe cascade nuclear refrigerator. The specimen was as-sembled in the same manner as the silver sample (seeFig. 15). To increase rigidity, rhodium foils were shapedin the form of the letter ‘‘U.’’ Owing to the poor thermalconductivity of Rh, the sample was connected to thecross-shaped thermal link via intermediate silver strips.

The initial polarization achieved in the rhodium spinassembly before demagnetization was limited, as in sil-ver, by the long spin-lattice relaxation time of Rh, t1514 h at 200 mK. Two mu-metal tubes, with a shieldingfactor of 400, were installed to reduce the remanent fieldof the demagnetization solenoid at the site of thesample. To reach negative temperatures, adiabatic de-magnetization of the specimen was stopped at 400 mTand a population inversion was made before continuingthe field sweep. As for silver (see Sec. IV.G), the inver-sion efficiency, 85–60%, depended on the initial polar-ization before the field flip. The effect was more dra-matic in rhodium. As is shown by Fig. 83, thepolarization after the field flip, as a function of the initialpolarization, reaches a maximum around pi'0.6. Thisbehavior remains a puzzle; it cannot be explained, forexample, by considering thermal mixing between theZeeman and the interaction reservoirs. For T,0, there-fore, the experimentally achievable polarizations werelimited to upu<0.60. Inversion fields between 100 mT and2 mT were tried, but the efficiency did not change much.

Examples of the measured dynamic susceptibilityx(f)5x8(f)2ix9(f), recorded by sweeping the fre-quency f of the excitation field, are displayed in Fig. 84,both when T.0 and when T,0. At negative tempera-tures x9,0, which again is a sign of energy emission.The static susceptibility x8(0)5(2/p)*(x9/f)df

Page 72: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

72 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

and polarization p5A*x9(f)df could then be calcu-lated. The integration was performed between 0 and 150Hz, and the proportionality constant A was determinedfrom experiments around 1 mK where the initial high-temperature polarization could be calculated fromp5tanh(mB/kBTe), using the pulsed Pt-NMR thermom-eter to determine T5Te . The platinum scale, in turn,was calibrated against the superconducting transitionpoint, 22.6 mK, of beryllium. In the ultralow-temperature region the temperature of the Rh nucleiwas again found by using the second law of thermody-namics, T5DQ/DS . For more details, see Sec. IV.

The local field B loc , caused by spin-spin interactions,was carefully determined in these experiments. B loc was

FIG. 83. Loss of polarization in rhodium during a field rever-sal: pf and pi are the final and initial polarizations, respec-tively. The ideal case is depicted by the solid line, while thedashed line shows the calculated result (see text). From Vuo-rinen, Hakonen, Yao, and Lounasmaa (1994).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

obtained by fitting the measured field dependence of thelongitudinal susceptibility xL to the theoretical expres-sion xL}(11B2/B loc

2 )23/2 (see Sec. IV.C). This equationhad worked reliably for calculating local fields in copperand silver (Ehnholm et al., 1980; Hakonen, Yin, andNummila, 1991). When applied to the data on rhodium,the equation yields B loc=(3463) mT, which is very closeto the result 35 mT found for silver, but much smallerthan B loc50.36 mT observed for copper.

Figure 85 displays the measured susceptibility of Rhas a function of polarization. For small upu, x8(0)follows the high-T approximation x8(0)5x0 /@12 (R1L2Dz)x0], where x05Ap5C/T is equivalent

to the Curie-law susceptibility of noninteracting spins(see Sec. IV.D). The data points were fitted to the theo-retical expression in the range 20.1,p,0.2; this yieldedR521.360.2 and the dash-dotted curve illustrated inFig. 85. Since R is on the order of one, the dipolar in-teractions may influence the ordered spin structures sub-stantially in rhodium.

One of the interesting features of the data in Fig. 85 isthat, at T,0, the susceptibility is a nearly linear functionof polarization all the way down to p5 –0.6. No satura-tion of x8(0) was observed as in silver at pc5 –0.49.

The absolute value of the inverse static susceptibilityof Rh nuclei as a function of uTu is shown in Fig. 86.The solid line represents the antiferromagnetic Curie-Weiss law, x5C/(T2uA), with C51.3 nK anduA5C(R1L2Dz)5 –1.4 nK, obtained at positive tem-peratures from the low-polarization data. At T,0, theferromagnetic dependence is displayed by the dashedline. At low temperatures the Curie-Weiss approxima-tion is known to deviate, especially when I5 1

2, from themore accurate results based on high-T series expansions(De Jongh and Miedema, 1974). For negative tempera-tures (filled circles), the measured data show a crossover

FIG. 84. NMR absorption x9 and dispersionx8 curves of rhodium nuclei measured in zeromagnetic field. Initial polarizations for T.0:h , p50.51; s , p50.32; n , p50.15. Initialpolarization for T,0: j , p520.51; d ,p520.35; m, p520.17. The solid curves areleast-squares fits to Lorentzian line shapes(see legend of Fig. 64), applied simulta-neously to the absorption and dispersioncurves. From Hakonen, Vuorinen, and Marti-kainen (1993).

Page 73: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

73A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

from ferro- to antiferromagnetic behavior around –5 nK.This indicates that the energy of the nuclear-spin assem-bly in rhodium is both minimized and maximized by an-tiferromagnetic order.

The data at T.0 extend to 280 pK and at T,0 to–750 pK. Both temperature records are roughly a factorof 2 closer to the absolute zero than the correspondingtemperatures reached during the measurements on sil-ver (see Sec. VIII). Phase transitions, however, were notseen in rhodium, even though the experimentally pro-duced polarizations of Rh nuclei, p50.83 and p

FIG. 85. Static susceptibility x8(0) vs polarization p ofrhodium nuclear spins in zero field for two specimens with 14and 6 ppm of effective iron impurities (open and filled circles,respectively). The dashed line is the Curie susceptibilityx051.16 p . The mean-field prediction x8(0)5x0 /@12(R1L2Dz)x0# [see Eq. (34)] is illustrated by the dash-dotted curve for R5 –1.35 and by the solid curve for R5 –1.0, using Dz50.05. Modified from Hakonen, Vuorinen,and Martikainen (1993).

FIG. 86. Absolute value of the inverse static susceptibilityu1/x8(0)u vs absolute value of temperature for rhodium nuclei:s , measured at T.0; d , measured at T,0. The straight lineswere calculated, using the Curie-Weiss law, from experimentaldata at small polarizations. The error bars denote the 20%uncertainty in the measurements of temperature. From Ha-konen, Vuorinen, and Martikainen (1993).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

5–0.60 at T.0 and at T,0, respectively, were higherthan those required for ordering in silver. At positivetemperatures, in particular, a careful search for orderingwas made. Disappointingly, no plateaus in the x(t)scans, characteristic of antiferromagnetic order (see Sec.VIII.C), were observed in experiments at 0, 19, and 36mT. The initial polarizations were 6, 14, and 15 percent-age units larger than those required for antiferromag-netic ordering in silver. The mean frequency of the ab-sorption spectrum did not show any kinks either.

B. Exchange constants and ordered spin structures

The susceptibility data can be used to extract thenearest- and next-nearest-neighbor exchange coeffi-cients J1 and J2 since the measured R yields ( jJ ij andB loc gives ( jJ ij

2 in Rh [see Eqs. (66) and (30a)]. This is,of course, possible only if interactions beyond second-nearest neighbors are neglected. The values obtainedfrom experimental results are J1 /h5 –1763 Hz andJ2 /h51063 Hz (Hakonen, Vuorinen, and Martikainen,1993). There is also another mathematical solution forthe exchange constants that is, however, unlikely be-cause it would require uJ2u@uJ1u.

It is interesting to know the ordered spin structuresthat correspond to these exchange constants. Mean-fieldcalculations, to be discussed in Sec. XV.B.7, have beenemployed to predict regions of different types of mag-netic ordering in the J1 vs J2 plane (Smart, 1966). Theresults are illustrated in Fig. 87. In an fcc lattice, thepossible ordered structures include the ferromagneticphase and three different antiferromagnetic states,known as structures of type I, II, and III. The magneticunit cells of these antiferromagnetic states are illustrated

FIG. 87. Phase diagram for magnetic ordering in rhodiummetal in the J1 vs J2 plane as predicted by the mean-fieldtheory (Smart, 1966). Regions of type-I, type-II, and type-IIIantiferromagnetic order (see Fig. 110) are depicted by AF-I,AF-II, and AF-III, respectively, while F refers to the ferromag-netic phase. Locations for Rh (circles) and Ag (triangles) aregiven; the open and filled symbols refer to positive and nega-tive spin temperatures, respectively. The ratio J1 /J2 is approxi-mately the same in Cu and in Ag, but the exchange constantsin Cu are larger by an order of magnitude (see Table IV).

Page 74: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

74 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

in Fig. 110. Rhodium lies well inside the type-I region atT.0. At T,0, the signs of the J’s are effectively re-versed. The corresponding point in Fig. 87 is located inthe ferromagnetic sector but rather close to the type-IIantiferromagnetic border. Neglecting the dipolar inter-action, the mean-field estimates of the Neel tempera-tures for the type-I and type-II orders are 1.5 nK and–0.7 nK, respectively (Vuorinen, Hakonen, Yao, andLounasmaa, 1995).

At T,0, description of rhodium using the J1,J2model fails since the experiments of Hakonen, Vuo-rinen, and Martikainen (1993) indicate antiferromag-netism rather than ferromagnetism. Inclusion of the di-polar interaction cannot explain the discrepancy, sincethis would just further stabilize the ferromagnetic state.It also seems unlikely that the discrepancy is caused bythe crudeness of the mean-field approximation. In fact,when thermal fluctuations are included, the phaseboundary between the ferromagnetic and type-II statesmoves further away from the position corresponding tothe observed exchange constants of rhodium (Heinilaand Oja, 1993b). A quite possible explanation is thatinteractions between third-nearest neighbors play a sig-nificant role, since J2 is only slightly smaller than J1.Pseudodipolar interactions could also be important indeciding between ferromagnetism and type-II antiferro-magnetism.

As another, at least partial explanation of the behav-ior of Rh at negative temperatures, one can argue thatthe polarization range 20.1,p,0.2, which Hakonen,Vuorinen, and Martikainen (1993) used to fit x8(0)against the high-T approximation to obtain R , was toonarrow (see Fig. 85). At small upu, only the simple Curiebehavior was observed and the fit was not sensitive toR . On the other hand, the high-T approximation isstrictly valid only at small upu. It is clear, however, that ifthe fit had been made in a larger polarization interval,the overall result would look much better since thebending of the best-fitting theoretical curve would de-crease at large upu. The mean-field curve with R521.0agrees with the data quite well, as is shown by the solidcurve in Fig. 85. Adopting this R and the measuredB loc we obtain J1 /h5 –16 Hz and Jnnn /h513 Hz. Thecorresponding point still falls within the ferromagneticregion in the J1J2 plane, but now the boundary to thetype-II state is within the error bar. Yet another indica-tion that the originally quoted R521.3 is too large inabsolute value comes from the susceptibility comparison(see Table VI in Sec. XV.B.5). The absolute values ofthe susceptibilities in type-I and type-II structures, ascalculated for the above J1 and J2, are about 30%smaller than the largest measured susceptibilities atT.0 and T,0, respectively. These discrepancies wouldalso decrease if uRu were reduced.

In any case, rhodium lies close to the border betweenthe ferromagnetic and the type-II antiferromagnetic re-gions at T,0. This should result in strong fluctuationsthat lower the ordering temperature from the mean-fieldestimate. It is therefore not surprising that ordering wasnot seen at negative temperatures. It remains puzzling,

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

however, why the ordered state was not observed atT.0 either. When the mean-field prediction for the or-dering temperature, TN

MF=1.5 nK, is corrected for fluc-tuations according to the high-temperature expansions(Pirnie et al., 1966), one obtains a theoretical TN around1.0 nK, much higher than the lowest temperature of 0.28nK reached in the experiments. Perhaps rhodium exhib-its behavior similar to that of silver (so far unexplained):The measured ordering temperatures of silver nucleiwere smaller than the theoretical predictions by a factorof 3 both at T.0 and at T,0 (see Sec. VIII.C.5).

C. Spin-lattice relaxation in rhodium

The acceleration in the spin-lattice relaxation, causedby electronic magnetic impurities, has not been investi-gated much in the microkelvin range and below, in spiteof the significance of t1 for reaching the lowest nucleartemperatures. However, the Helsinki group recentlymade measurements on two rhodium specimens (Ha-konen, Vuorinen, and Martikainen, 1994). Surprisingly,the data show that in weak magnetic fields there is aclear difference in t1 at T.0 and T,0. The observedbehavior is consistent with electrons scattering from ironimpurities, provided that the scattering rate has a contri-bution proportional to the inverse of the nuclear-spintemperature.

The spin-lattice relaxation time is defined (see Sec.II.B) by the relationship

d~1/T !/dt52~1/t1!~1/T21/Te!. (52)

Since Te@T and p}1/T [see Eq. (32)], one obtainsd lnp/dt52(1/t1). The cooling procedure and thesamples in the relaxation experiments were the same asused in earlier studies on rhodium (see Sec. X.A). Twomethods were employed in the experiments. Between 40and 400 mT, a fixed-field technique was applied by moni-toring, at successive times, the area under the NMR ab-sorption or, at T,0, emission signal x9; this againyielded the polarization according to the relationp5A*x9(f)df . The spin-lattice relaxation time wasfound by plotting the measured points on a log p vs tgraph (see Fig. 88). The straight line, least-squares fittedto the data obtained during the second half of the ex-periment, illustrates a constant relaxation rate, while thesteeper slope at the beginning of the run indicates amuch shorter t1.

In order to compare the relaxation rates at T.0 andT,0 directly, population inversions were made at regu-lar time intervals while recording p . Data at the begin-ning of such an experiment, performed at 40 mT, areillustrated in Fig. 89. The results show clearly that t1 islonger at T,0 than at T.0, even though the spin-lattice relaxation slows down notably with time whenT.0. This unexpected effect is much smaller in highfields but, when B50, the measured t1 is almost a factorof 2 longer at negative temperatures.

The second method for measuring t1, a field-cyclingsequence, was used in small fields, B<100 mT, and alsoat 1 mT ,B,1 T. First, 5–7 NMR spectra were re-

Page 75: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

75A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

corded at 400 mT to obtain the initial polarization. Next,B was swept, in 5–15 s, to the desired low field and thenkept constant for a time Dt51 –90 min, after which thefield was returned back to 400 mT and 5–7 NMR spectrawere again recorded. The decay of polarization duringDt , from which the relaxation at B5400 mT was re-moved, was then employed to calculate t1. Before thenext experimental cycle was started, a population inver-sion was made so that data at T.0 and T,0 were againmeasured consecutively.

The experimental results on the sample with 6 ppm ofiron impurities are summarized in Fig. 90, which displays

FIG. 88. Polarization p of rhodium nuclei as a function of timeat T.0 and B5100 mT. A constant spin-lattice relaxationtime is characterized in this semilogarithmic plot by thestraight line, which is a least-squares fit to the latter half of thedata points. The polarization dependence of t1 is quite obviousat the beginning of the experiment. From Vuorinen, Hakonen,Yao, and Lounasmaa (1995).

FIG. 89. Polarization upu of rhodium spins as a function of timemeasured in a magnetic field of 40 mT: s, at T.0; d , atT,0. The straight line is a least-squares fit to the data atT,0. During the blank sections the polarization was checkedat 400 mT, and the spin populations were inverted by rapidfield reversals. From Hakonen, Vuorinen, and Martikainen(1994).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

the spin-lattice relaxation times measured in magneticfields B50, 50, 100, and 400 mT at the electronic tem-perature Te5120 mK. The salient features of the dataare that t1 clearly decreases with increasing upu and de-creasing B . In addition, there is a clear asymmetry int1 between T.0 and T,0, especially in low fields.

If the usual Korringa’s law is valid for the specimenunder study, the spin-lattice relaxation time should havebeen t15k0 /Te=23 h=8.3·104 s for the data in Fig. 90 infields larger than B loc534 mT. In fields lower thanB loc , t1 should always be shorter (see Sec. II.B), butonly by a factor between 2 and 3. Instead, the zero-fielddata show reductions by a factor of about 50 from thevalue predicted by high-field Korringa constant k0510sK. It seems, therefore, difficult to explain the observeddifferences in t1 at T.0 and at T,0 by intrinsic prop-erties of rhodium.

The magnetic-field dependence of k has been plottedat a constant Te for the two Rh specimens in Fig. 91.The Korringa constant begins to decrease in fieldssmaller than 10 mT for both samples. The relaxation isfaster, by a factor of 5, in the sample with 14 ppm of Feimpurities than in the sample with 6 ppm of Fe. Thereduction of k in Rh was therefore attributed to ironimpurities. Similar results had been observed previouslyin copper (Huiku, Loponen, et al., 1984). The data onRh also show that the effect is enhanced when the elec-tronic temperature is lowered.

The scattering rate t1imp(B), caused by impurities,

can be related to the observed t1 by the equationt1

215Te /k01@t1imp(B ,T)#21. According to phenomeno-

logical arguments, the impurity scattering rate, affectedby the susceptibility-dependent field distribution arounda magnetic center (Hakonen, Vuorinen, andMartikainen, 1994), is of the form @t1

imp(B ,T)#21

FIG. 90. Spin-lattice relaxation time t1 of a rhodium samplewith 6 ppm of iron impurities as a function of the high-fieldpolarization p of the Rh nuclei; d , measured at B50; j , n ,B550 mT; s , B5100 mT; h , B5400 mT. The open and filledsymbols refer to measurements performed with the fixed-fieldand the field-cycling techniques, respectively. From Hakonen,Vuorinen, and Martikainen (1994).

Page 76: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

76 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

5@tp50(B)#21@11jx(B)# , where tp50(B) is the relax-ation time in a magnetic field B at zero spin polarizationand j is a constant close to unity. In Fig. 92, data pointshave been compared with these arguments for j50.5using the measured x . Within experimental uncertain-ties, the calculated curves agree with the measured re-sults. It can then be concluded that the origin of theanomalous behavior of t1 is scattering from isolated Feimpurities, modified by the large susceptibility of Rh nu-clei. The change in t1

imp(B ,T) is proportional to 1/T .One property that distinguishes Rh from Cu and Ag is

the temperature dependence of electrical resistivity

FIG. 91. Field dependence of Korringa’s constant for a Rhsample with 6 ppm of iron impurities: n , at Te50.12 mK;s , at Te50.2 mK; h , at Te51.2 mK; d , a 14 ppm sample atTe50.2 mK. The nuclear-spin temperature is positive in allcases. The dashed curves are only guides for the eye. The solidline in the high-field region depicts the behavior of platinum(Roshen and Saam, 1980, 1982). From Vuorinen, Hakonen,Yao, and Lounasmaa (1995).

FIG. 92. Spin-lattice relaxation times of Fig. 90, scaled by thepolarization dependence of t1 at B5400 mT. The fittedstraight lines have been drawn with equal slope. For furtherexplanation, see text. From Hakonen, Vuorinen, and Marti-kainen (1994).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

caused by magnetic impurities. RhFe belongs to the so-called Coles alloys (Rivier and Zlatic, 1972), which dis-play a positive temperature coefficient in their low-temperature resistivity. In these materials (IrFe, PtFe,PdFe, . . . ), both the host and the impurity are transitionmetals, whereas in Kondo alloys (AuMn, AlMn, . . . ),which exhibit a resistance minimum, the host and theimpurity have different electronic structures. However,according to theory, both classes of alloys should displayvery similar universal behavior that can be understoodin terms of electrons scattered from localized spin fluc-tuations around the impurity (Roshen and Saam, 1980,1982). Such a description is not successful for RhFe; thedecrease in k predicted by theory is four orders of mag-nitude too large.

Spin-lattice relaxation measurements on copper havebeen made by Smeibidl, Schroder-Smeibidl, and Pobell(1994) between 60 mK and 10 mK on three specimenshaving RRR values of 260, 960, and 1300, respectively.No deviations from the Korringa law were observed.

XI. NMR STUDIES IN THE PARAMAGNETIC PHASE

In this section we describe some NMR measurementson highly polarized nuclear spins. Although these ex-periments were performed in the paramagnetic statethey also reveal features important for nuclear ordering.There are several ‘‘high-temperature’’ NMR techniquesthat yield information on the absolute magnitude ofspin-spin interactions but not on the sign; for a discus-sion of these methods we refer the reader to Oja (1987).

We first review measurements of cross relaxation be-tween the Zeeman temperatures of 107Ag and 109Ag(Oja, Annila, and Takano, 1990). This work elucidatesthe important thermal mixing process in which the coldZeeman reservoir cools the interaction reservoir afterdemagnetization to a low field (see Sec. II.B). Next, wediscuss the so-called ‘‘suppression-enhancement’’ of iso-topic NMR lines. This effect has been observed in two-isotope systems such as those in copper (Ekstrom et al.,1979) and in silver (Oja, Annila, and Takano, 1990; Ha-konen, Nummila, and Vuorinen, 1992). An analysis ofthe data yields the overall strength of the exchange in-teraction, i.e., the parameter R defined by Eq. (6). Theresult can be directly compared with that obtained fromthe Curie-Weiss fit to the static susceptibility vs tem-perature curve or with electronic band-structure calcula-tions. Finally, we discuss the observation and analysis ofthe second-harmonic NMR absorption peak in low fieldsfor highly polarized copper spins. The position of thepeak also yields R . This technique can be used in single-isotope spin systems as well.

A. Cross relaxation between 107Ag and 109Ag

Cross relaxation in spin systems has been investigatedextensively since the NMR experiment of Abragam andProctor (1958) and the studies by Bloembergen and Per-shan and co-workers (Bloembergen, Shapiro, Pershan,and Artman, 1959; Pershan, 1960). The majority of later

Page 77: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

77A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

investigations have been on rotating-frame Zeeman sys-tems using the spin-locked NMR technique (Schmid,1973). The studies on silver by Oja, Annila, and Takano(1990) involved a conceptually simpler case, cross relax-ation between two laboratory-frame Zeeman systems.As their data extended to many orders of magnitudelower temperatures than before, it was possible, for thefirst time, to investigate the polarization dependence ofthe cross-relaxation time.

Natural silver is a mixture: 51.8% of 107Ag and 48.2%of 109Ag. The two isotopes have a 13% difference intheir gyromagnetic ratios: g107/2p=1.73 MHz/T andg109/2p=1.98 MHz/T. This suggests that thermodynami-cally the system can be described using four heat reser-voirs, each of them having its own temperature: T107 andT109 for the two Zeeman reservoirs, Tss for the commonspin-spin interaction reservoir through which T107 andT109 communicate, and Te for the lattice and the con-duction electrons.

The reason why one must assign separate Zeemantemperatures for the different spin populations is rathersimple. Consider two nearby silver spins, one of the 107isotope and the other of the 109 isotope. Energy is trans-ferred between the two spin populations when one ofthe spins is flipped (↓→↑) and the other is flopped(↑→↓). The overall Zeeman energy is not conserved inthe flip-flop process because the gyromagnetic ratios ofthe two spin species are different. The law of energyconservation requires then that an equal but oppositechange in energy occur in the interaction reservoir. Asthe spread of the spin-spin interaction energies is re-flected in the NMR linewidth, one might expect that therate at which T107 and T109 approach each other is pro-portional to the degree of overlap of the two NMR lines(Bloembergen, Shapiro, Pershan, and Artman, 1959).The higher the magnetic field, the smaller is the overlap,and the longer is the cross-relaxation time tx .

The data of Oja, Annila, and Takano (1990) showedthe expected increase of tx with B . It was remarkable,however, that the observed cross-relaxation times wereorders of magnitude longer than what one would haveexpected from the overlap of the NMR lines. The reasonunderlying the failure of the naive argument highlights afundamental aspect of spin-temperature theory: There isa hierarchy of relaxation processes and times in the sys-tem, associated with mutual flips of unlike spins, single-spin flips, and finally the t1 process, all with their own,well-defined (quasi-)equilibrium states.

1. Measurement of the cross-relaxation time tx

The specimen of Oja, Annila, and Takano (1990),consisting of 28 polycrystalline silver foils, was mountedas the second nuclear stage in the Helsinki refrigerator(see Sec. III.B). An rf SQUID, connected to an astaticpickup coil, detected the NMR signal. A linear excita-tion field of 40-nT amplitude was used. The measuringfrequencies were below 2 kHz, which guaranteed fullpenetration of the excitation field into the sample. Sincethe nuclei were practically isolated from the lattice inthese measurements (the spin-lattice relaxation times

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

were about 14 h), very slow cross-relaxation processescould be investigated. For other details of the experi-mental procedure, see Secs. III.B and III.E.

The final step of cooling consisted of lowering thefield on the silver specimen from 7.3 T to 0.93 mT. Thecontinuous-wave NMR signal, illustrated by curve (a) inFig. 93, was then measured in order to determine theinitial polarization. Next, the 109Ag NMR signal was al-most saturated by applying a strong audio-frequencyfield near the 109 resonance. The Zeeman temperatureT109 of the 109Ag spins (curve b in Fig. 93) thereby be-came much higher than T107 . The magnetic field wasthen lowered to a holding value of 0.77 mT for experi-mental convenience (curve c in Fig. 93). In this field,tx is practically infinite.

However, when the field was reduced to 0.22 mT, foronly a short time Dt=0.20 s, and then increased back to0.77 mT, a different spectrum was obtained: the twopeaks were now approximately equal (curve d). Thismeans that the 109 spins had cooled down and the 107spins had warmed up. Two more cycles and the spec-trum (curves e and f) began to look as it had in theoriginal equilibrium situation (curve a). Thus, by meansof the field-sweep sequence 0.77 mT → 0.20 mT → 0.77mT, ‘‘snapshots’’ of thermal equalization between thetwo spin species were obtained.

The cross-relaxation time was calculated from thechange in the NMR signals as a function of the cumula-tive period spent in the low field, at 0.20 mT. A singleexponential time constant described the changes in theintegrated areas under the 107Ag and the 109Ag peaks.For data in the highest fields, the intensities of the NMRsignals had to be corrected for decay caused by the spin-lattice relaxation described by t1. In the low-field mea-surements, B was changed quickly in order to ensurenegligible cross relaxation during the field sweep. To de-termine the dependency of tx on the spin temperature, asecond cross-relaxation measurement was usually madeimmediately afterwards by starting from the final state

FIG. 93. NMR spectra demonstrating cross relaxation in silver.For explanations, see text. From Oja, Annila, and Takano(1990).

Page 78: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

78 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

of the first run (curve f in Fig. 93): the 109Ag spins wereagain heated and the snapshot procedure was repeated.

The data of Oja, Annila, and Takano (1990) cantherefore be divided into two sets. The high-polarizationresults, obtained in the first cross-relaxation experimentsduring run (a), correspond to the following initial condi-tions after the 109Ag signal was almost saturated:p107=0.41–0.61 and p10950.16 –0.31; T107545 –73 nKand T1095110–230 nK. The low-polarization data, col-lected during the second series of experiments, relate toinitial conditions p10750.11–0.20 and p10950.04–0.11.The measured values of tx , as a function of B2, arepresented in Fig. 94; the data cover five decades. In thelowest fields tx't2510 ms. When B>0.15 mT, logtx isapproximately linear in B2, although the increase in txbecomes less steep at B' 0.35 mT. Comparison of thehigh-p and low-p data shows that the cross-relaxation

FIG. 94. Time constant tx for cross relaxation in silver as afunction of the external magnetic field squared: d , high polar-ization; s , low polarization. The solid lines are theoreticalcurves for tx at low p , using various values for the antiferro-magnetic nearest-neighbor RK interaction. (a) J/h520 Hz,(b) J/h531 Hz, (c) J/h540 Hz. Inset: The ratior5tx(low p)/tx(high p) vs B ; the curve is only to guide theeye. From Oja, Annila, and Takano (1990).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

time decreases with polarization. However, the ratiotx(low p)/tx(high p) clearly increases with B , as is il-lustrated in the inset of Fig. 94.

2. Theoretical description

The quantum-statistical theory of cross relaxation wasdeveloped by Provotorov (1962). To describe cross re-laxation in a two-isotope system, such as the naturalmixtures of silver and copper, it is useful to present thesystem in terms of the four nuclear-spin heat reservoirsillustrated in Fig. 95 (Goldman, 1970): the average Zee-man energy HZ

av , the differential Zeeman energy HZdiff ,

the interaction reservoir Hss8 , and the conduction-electron reservoir. The average and the differential Zee-man energies are defined by

HZav52\gavB(

iI i

z, (53)

HZdiff52\~ga2gb!B(

i~xbIa ,i

z 2xaIb ,iz !, (54)

where gav5xaga1xbgb and the concentrations of thetwo spin species a and b are xa and xb , respectively.The direction of the external field is denoted by z . Thetotal Zeeman energy HZ for the two spin species equalsthe sum HZ

av1HZdiff . The usefulness of this decomposi-

tion becomes evident when we consider cross relaxationcaused by mutual spin flips.

The interaction reservoir Hss8 consists of those terms inthe Hamiltonian that commute with HZ . The conservingand nonconserving terms can be distinguished by intro-ducing the conventional notation using raising and low-ering operators, defined as Ii

15Iix1iI i

y and Ii25Ii

x2iI iy

respectively (see, for example, Van Vleck, 1948). Thefull spin-spin interaction Hss5HD1HRK [see Eqs. (2)and (3)] can then be written as

Hss5A1B1C1D1E1F, (55)

where

A5(i,j

aijI izIj

z, (56a)

FIG. 95. Heat reservoirs in a two-isotope system: 107Ag and 109Ag nuclei and conduction electrons at low temperatures. Thereservoirs are for the average Zeeman energy HZ

av , the differential Zeeman energy HZdiff , the interaction reservoir Hss8 , and the

conduction-electron reservoir at Te . See text for details.

Page 79: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

79A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

B5(i,j

bij~Ii1Ij

21Ii2Ij

1!, (56b)

C5(i,j

c ij~Ii1Ij

z1IizIj

1!, (56c)

D5C†, (56d)

E5(i,j

e ijI i1Ij

1, (56e)

F5E†. (56f)

Here O† denotes the Hermitian operator of O. The no-tation ( i,j implies that pairs (i ,j) are to be counted onlyonce. The various coefficients appearing above are

aij52Jij1\2g ig jr ij23@123~cosu ij!

2# , (57a)

bij52 12 Jij1

14 \2g ig jr ij

23@3~cosu ij!221# , (57b)

cij52 32 \2g ig jr ij

23sinu ijcosu ijexp~2if ij!, (57c)

eij52 34 \2g ig jr ij

23~sinu ij!2exp~22if ij!, (57d)

where u ij and f ij are the azimuthal and polar angles ofrij with respect to the z axis.

The expression for the truncated spin-spin interactionHss8 is

Hss8 5A1Baa1Bbb, (58)

where the superscripts aa and bb denote the species ofthe interacting spins i and j . Therefore mutual flips oflike spins are included in Hss8 , but those for unlike spinsare not since the latter do not conserve the Zeemanenergy, i.e., @Bab1Bba,HZ#Þ0. Neither do single flips(C`D) or double flips (E`F) conserve HZ .

The operators HZav , HZ

diff , and Hss8 are mutually or-thogonal, i.e., they commute with each other. Thus eachof them can be viewed as a heat reservoir having its owntemperature. When spin-lattice relaxation can be ne-glected, these reservoirs approach a common tempera-ture only through spin-spin interactions not contained inHss8 . In an external field, the most frequent of these pro-cesses are mutual flips of unlike spins, since one suchprocess changes the Zeeman energy only by(ga2gb)B . These processes bring HZ

diff and Hss8 intothermal equilibrium with each other. The temperatureof the average Zeeman reservoir is not affected, how-ever, since @Bab1Bba,HZ

av#50. This explains the failureof the overlap argument of isotopic NMR lines in ac-counting for the cross-relaxation time constant: mutualflips of unlike spins do not bring about thermal equilib-rium (Goldman, 1970; Rodak, 1971).

Therefore thermal equilibrium is obtained via slowerprocesses involving single spin flips, described by C`Din Eq. (55). These terms are caused by the dipolar inter-action alone, whereas mutual spin flips depend also onthe exchange interaction. During a single spin-flip pro-cess, the reservoirHZ

diff1Hss8 1Bab1Bba can be treated asa single system in internal thermal equilibrium. The factthat there are two different spin species plays then only

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

a minor role, since the process is essentially the same asthermal mixing, which equalizes the temperatures of theinteraction and Zeeman reservoirs in systems with onlyone kind of spin (see Sec. II.B and Fig. 6). The presenceof two spin species is important, however, for the detec-tion of cross relaxation, since the process shows up asopposite changes in the isotopic Zeeman temperatures,which can be measured by continuous-wave NMR tech-niques, and because the initial nonequilibrium state canbe realized by simply saturating one of the resonances.

3. Extraction of the exchange constant

Cross relaxation offers, in principle, a very sensitiveprobe for determining the strength of the exchangeinteraction. Using Provotorov’s (1962) theory, tx can becomputed for given spin-spin interactions, although thecalculation is somewhat complicated. Oja, Annila, andTakano (1990) computed tx numerically by taking intoaccount dipolar interactions up to the sixth shell ofneighbors and by assuming different strengths for thenearest-neighbor RK interaction 2JIi•Ij . More distantJ’s were neglected, since their values are small. Therelevant spin-correlation function was assumed tohave the Gaussian form (Goldman, 1970; Demco,Tegenfeldt, and Waugh, 1975). It then follow, approxi-mately, that tx5t0exp(B2/b2), where the constant b ison the order of B loc and t0 is on the order of the spin-spin relaxation time t2 (Goldman, 1970). The rather lin-ear behavior of the measured tx in a logtx vs B2

plot supports this assumption (see Fig. 94), althoughin high fields the increase in tx with B was not sosteep. For uJu/h.20 Hz, it was found that b2

'11J2/(\2gav2 )1xaxb(ga2gb)2B2/gav

2 . The best fit tothe experimental data on silver was obtained with anantiferromagnetic J/h52(3162) Hz; the agreementbetween theoretical and experimental values of tx israther good.

Poitrenaud and Winter (1964) deduced from a line-width analysis of their NMR data at high temperaturesuJu/h5(26.561.5) Hz, but they were unable to extractthe sign of J . As long as uJu is much larger than thenearest-neighbor dipolar interaction, tx is also rather in-sensitive to the sign of J . The relative changes in theequilibrium intensities of the 107Ag and 109Ag signals asa function of polarization show clearly, however, that Jis antiferromagnetic. This so-called ‘‘suppression en-hancement’’ will be discussed in detail in Sec. XI.B. Thiseffect, too, yields J consistent with the tx analysis. Inaddition, J/h52(3162) Hz is in good agreement withband-structure calculations (Harmon et al., 1992) andsusceptibility measurements (Hakonen, Yin, and Lou-nasmaa, 1990).

Buishvili and Fokina (1994) have recently investigatedthe dependence of the cross relaxation on polarization.Their theoretical calculations show that, in sufficientlylow fields, tx decreases with increasing p and that theratio tx(low p)/tx(high p) increases with B , as was ob-served in the measurements (see Fig. 94).

Page 80: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

80 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Knowledge of tx is crucial in the interpretation oftime-dependent phenomena associated with spontane-ous nuclear ordering. In fields below 0.1 mT, wherenuclear ordering has been observed (see Fig. 73),tx't2 and cross relaxation is fast in comparison withtypical demagnetization and remagnetization rates usedin the nuclear-ordering experiments. Assuming that thesame holds for copper, slowness of cross relaxation can-not be the source of nonadiabaticity (see Secs. V.C andVII.F.5), contrary to a proposal by Lindgard (1988a).This is important to know, since thermometry in theseexperiments (see Sec. IV) relies on adiabatic fieldchanges.

B. Polarization-induced suppression and enhancementof isotopic NMR lines

1. Copper

The suppression-enhancement effect was discoveredby Ekstrom et al. (1979) when they made NMR mea-surements of highly polarized copper spins. The effect isillustrated in Fig. 96. It shows the NMR absorptionx9(f), measured using the field-sweep technique at thefrequency f5183 kHz. Peaks for the two isotopes 63Cuand 65Cu are clearly separate at all polarizations. Thepositions of the peaks correspond approximately tothe gyromagnetic ratios g63/2p511.3 MHz/T andg65/2p512.1 MHz/T. The intensities of the two reso-nances are proportional to the abundancies of the iso-topes, x6350.69 and x6550.31, only at small polariza-tions. At high p , there is a clear discrepancy between theexpected and observed intensity ratios. When p50.9,the signal from the less abundant isotope 65Cu is muchlarger than the signal from 63Cu. The intensity ratio,determined from the areas of the peaks, is 1.7 and de-creases towards 0.5 in the low-polarization limit.

Ekstrom et al. (1979) explained this effect by the useof an internal-field model. The local ac field on 63Cuspins consists of the external excitation field B1 and theinternal fields due to exchange and dipolar interaction

FIG. 96. NMR absorption x9(f) of copper at the frequencyf5183 kHz, measured by sweeping the external magnetic field.Nuclear-spin polarizations p have been indicated for eachcurve. From Ekstrom et al. (1979).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

with 65Cu, and vice versa. The latter is proportional tothe magnetization. The effect of the exchange interac-tion is to replace the magnetic field B05Bextz1B1x byBi* 5B01m0RMj , acting on the isotope i . The ac field ismodified, in turn, by the factor @12Rx

j(v)#21, and far

from the resonance frequency v j* 5g jBjz* one may ap-proximate by noting that x

j}(v j* 2v)21. Thus for

R,0, the excitation at the higher resonance frequency isenhanced by the other isotope, as is shown by the datain Fig. 96.

This model was employed to determine the exchangeconstant R . The observed line shapes could be repro-duced by assuming R520.4360.04. This value is some-what different from R520.39, obtained from suscepti-bility measurements (see Sec. V.A). The theoreticallycalculated exchange interactions yield a lower R as well(see Table IV in Sec. XV.A). It seems also that the valueuRu50.43 is 10–20 % too large to be consistent with theNMR linewidth measurements of Andrew et al. (1971).

2. Silver

Silver is another two-isotope system, consisting of107Ag and 109Ag, in which the suppression-enhancementeffect has been investigated. Measurements at T.0were performed by Oja, Annila, and Takano (1990) andat T,0 by Hakonen, Nummila, and Vuorinen (1992).Figure 97 shows absorption spectra at both positive andnegative polarizations upu'70%. There is an interestingqualitative difference between the spectra. At T.0, thepeak at the higher frequency was enhanced with increas-ing polarization. This was caused by the antiferromag-netic exchange interaction, as in copper. At T,0, how-

FIG. 97. NMR absorption x9(f) of silver, scaled to the exter-nal field B50.8 mT, at positive and negative polarizationsupu'0.7. The curve at T.0 was obtained from the data of Oja,Annila, and Takano (unpublished) at B50.774 mT andp50.70, after scaling their frequency axis by 1.034. The spec-trum for T,0 is from measurements of Hakonen, Nummila,and Vuorinen (1992) at B50.822 mT and p520.68, scaleddown in frequency by the multiplier 0.973.

Page 81: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

81A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

ever, the opposite behavior was observed, as if theinteractions were ferromagnetic. This can be understoodin terms of internal fields that depend on the exchangeinteraction through the product pR . Therefore, by re-versing the sign of either polarization or the exchangeinteraction, one obtains the same behavior.

The other characteristics of the spectra in Fig. 97 in-clude an overall down shift of the two resonances atT.0, while an up shift is observed at T,0. Finally, in afashion similar to the repulsion of any coupled energylevels in second-order perturbation theory, increasingpolarization makes the two NMR peaks repel each otherat positive and negative temperatures.

Hakonen, Nummila, and Vuorinen (1992) presenteddata on the intensity ratio of the two absorption lines insilver for polarizations in the range 0 – 20.70. In addi-tion, the separation of the peak frequencies was mea-sured. At positive temperatures, the ratio of the intensi-ties for the 107Ag and 109Ag lines had been measured byOja, Annila, and Takano (1991). These authors used thesuppression-enhancement effect for matching the spin-polarization scales of their two experimental setups.

The internal-field model of Ekstrom et al. (1979)was applied, quite successfully, to describe the data bothat T.0 and T,0. The best fit to the experimental re-sults was obtained with the exchange parameterR=22.3 – 22.5, in good agreement with the values ob-tained from susceptibility measurements (Hakonen, Yin,and Lounasmaa, 1990), from cross-relaxation data (Oja,Annila, and Takano, 1990), from an analysis of theNMR linewidth (Poitrenaud and Winter, 1964), andfrom electronic-structure computations (Harmon et al.,1992).

3. Theoretical calculations

The internal-field model of Ekstrom et al. (1979) hasrecently been elaborated by Heinila and Oja (1994d).These authors give analytical expressions for the inten-sities and positions of the resonance peaks for spins in-teracting via exchange forces only. More importantly,they introduce a new method for accurately calculatingNMR line shapes for a system consisting of classicalspins. The technique is based on a combination ofMonte Carlo simulations and numerical solutions of themicroscopic equations of motion. The calculations wereused to establish the accuracy of the approach used byEkstrom et al. (1979). It was found that the model workswell, especially at intermediate polarizations when thelines are clearly separated. In certain cases, however,Heinila and Oja observed that the internal-field modelfails. These authors also investigated how the NMR re-sponse is affected by the random distribution of spinsover the lattice sites.

Eska (1989) has studied, by the use of numerical simu-lations, the response of a system with two isotopes tolarge NMR tipping pulses. His results show, for ex-ample, that the roles of suppressed and enhanced reso-nances are exchanged when the tipping angle ap-proaches p/2. It would be interesting to test thisprediction by extending NMR studies of highly polar-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

ized spins to large tipping angles; the experiments oncopper and silver were performed using continuous-wave NMR in the limit corresponding to small tippingangles.

The suppression-enhancement effect has also been in-vestigated theoretically by employing the perturbationapproach of Kubo and Tomita (1954; Oja, Annila, andTakano, 1988) and by the Green’s-function technique(Buishvili, Kostarov, and Fokina, 1994). These calcula-tions, as well as the simulations of Heinila and Oja(1994d), can also account for the fact that exchange in-teractions cause merging of the isotopic absorption lineswhen their Larmor frequencies are sufficiently close toeach other and the polarization is small.

C. Second-harmonic Larmor line

In their measurements of highly polarized copperspins, Ekstrom et al. (1979) observed an interestingNMR phenomenon related to the second-harmonic Lar-mor line, i.e., absorption near v2L52gB . The line isvery difficult to see in high external fields since, accord-ing to Anderson (1962), the intensity of the v2L reso-nance is approximately proportional to (11B2/b2)21

times the intensity of the vL peak, where b is on theorder of the local field. As B loc50.34 mT in copper, thev2L resonance can be observed in fields around 1 mT.This is illustrated by Fig. 98. The second harmonic isseen clearly in the data for the polarization p50.9 abovethe main peak. The position of the resonance is shiftedbelow the high-temperature value v2L/2p52gB . Atp50.15, the intensity of the second harmonic is low butsufficient to show that the resonance is located at ahigher frequency than when p50.9.

FIG. 98. NMR absorption x9 of copper in the external mag-netic field B51.1 mT. Nuclear-spin polarization is 0.9 in theupper frame and 0.15 in the lower. The vertical scale has beenexpanded by 4 in the lower frame. From Ekstrom et al. (1979).

Page 82: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

82 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Ekstrom et al. were able to explain the origin of theshift by analyzing the equations of motion for spins ro-tating at frequencies near the Larmor value and its sec-ond harmonic. It was found that coupling between thesemodes of rotation yielded the estimate R=20.4160.04for the strength of the exchange interaction, in goodagreement with R=20.4360.04 obtained from an analy-sis of the suppression-enhancement effect. The fact thatthe shift in the v2L line is definitely negative sets anupper limit of 20.33 for R .

Moyland et al. (1993) have extended the work of Ek-strom et al. (1979) by calculating the intensities and reso-nance frequencies for the vL and v2L lines in silver andgold, as well. They found that the intensities of thev2L lines decay quickly away from the field in whichvL and v2L resonance frequencies anticross. This iscaused by the large exchange interactions in these met-als. Detection of the second harmonics would thereforebe difficult. The authors conclude, however, that deter-mination of R through coupling of the vL and v2Lmodes may still be possible by observing a break in theLarmor mode at the anticrossing field.

XII. EXPERIMENTS ON THALIUM, SCANDIUM, AND AuIn2

In these metals, with k54.4 msK for Tl, 90 msK forSc, and 110 msK for AuIn 2, the spin-lattice relaxationtime t1 is so short that the conduction-electron andnuclear-spin temperatures are essentially equal. Thismeans that the exchange interaction is strong, probablyproducing nuclear order at a relatively high tempera-ture, and that thermal isolation of the nuclei from exter-nal sources of heat leaks is not possible relying on a longspin-lattice relaxation. Consequently, experiments on Tl,Sc, and AuIn 2 have been carried out in quite differentregimes than those on Cu, Ag, and Rh. Instead of nano-and picokelvin nuclear temperatures, one works in themicrokelvin range and with T5Te (Pobell, 1994).

In this section we shall first discuss briefly the experi-ments on thallium, in which NMR measurements haveshown interesting behavior at low temperatures (Eskaand Schuberth, 1987; Leib et al., 1995). We then present,in more detail, the data on scandium (Suzuki et al., 1994;Koike et al., 1995) and on AuIn 2 (Herrmannsdorfer andPobell, 1995; Herrmannsdorfer, Smeibidl et al., 1995).Magnetic susceptibility measurements on Sc show thatthis metal probably orders ferromagnetically at about100 mK. In the intermetallic compound AuIn 2 a ferro-magnetic transition has been discovered at 35 mK. Thisis the first unambiguous observation of spontaneousnuclear magnetic ordering in a non-hyperfine-enhancedmetal (see Sec. XIII) with thermal equilibrium betweenthe nuclei and the conduction electrons. We would liketo remind the reader, however, that nuclear-spin orderin Cu, Ag, and Rh is not influenced by the temperatureof conduction electrons. Besides, the spin temperatureof the gold nuclei may differ from Te owing to the muchlarger Korringa constant of Au in AuIn 2.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

A. Thallium

The NMR behavior (Eska and Schuberth, 1987) andnuclear specific heat (Schroder-Smeibidl et al., 1991) ofthallium have been investigated using two-stage nuclear-demagnetization refrigerators in Garching and inBayreuth. Among metallic elements, thallium has thestrongest known nucleus-electron coupling, which re-sults in the Korringa constant k5t1Te54.4 msK (Eskaet al., 1986). The absolute strength of exchange interac-tions has been determined by studying merging of theNMR absorption lines of 203Tl and 205Tl (Karimov andShchegolev, 1961). Assuming only nearest-neighbor ex-change interactions, uJu/h537.5 kHz. If the interactionis ferromagnetic, one obtains the mean-field estimateTC5 1

3I(I11)zJ/kB=5 mK using z512 and I5 12. Unfor-

tunately, thallium is toxic and readily oxidizes when it isexposed to air. This will severely limit the use of thismetal as a nuclear refrigerant.

Eska and Schuberth (1987) refrigerated their 4N-purethallium samples to temperatures as low as 70 mK. Achange in the NMR signal was observed for spin polar-izations exceeding 40%: the single line split into twopeaks with a nearly temperature- and field-independentseparation. This was, at the time, attributed to the onsetof nuclear-spin ordering. The intensity of the NMR line,above the assumed transition, was enhanced over thevalue expected from a Curie-law susceptibility.

Schroder-Smeibidl et al. (1991) later measured thenuclear heat capacity of a 0.79 mol thallium sample, 5N-pure, at 70 mK <T <20 mK and 20 mT <B<230 mT.The specific-heat data indicated simple nuclear para-magnetic behavior in the investigated temperature andfield regions. Therefore the NMR anomalies observedearlier by Eska and Schuberth (1987) cannot have re-sulted from phase transitions. Spontaneous nuclear or-dering in thallium metal must thus occur well below 70mK, as was originally expected.

NMR behavior of thallium has also been investigatedin several theoretical studies. Calculations based on thelinear-response theory have not been able to reproducethe observed NMR anomalies (Oja, Annila, and Ta-kano, 1988; Eska, 1989; Heinila and Oja, 1994d; Buish-vili, Kostarov, Fokina, unpublished). According to re-cent theoretical and experimental work by Eska and co-workers, the line splitting results from the combinationof two effects: nonlinear spin dynamics due to a largeNMR tipping angle and interference effects of the mag-netization gradients in the rf penetration depth (Baumlet al., 1994; Leib et al., 1995).

B. Scandium

Nuclear-spin order in scandium metal has been stud-ied in Tokyo and Kanazawa by Suzuki and co-workers(Suzuki et al., 1994; Koike et al., 1995). Behavior thatmight have been due to nuclear ferromagnetism was ob-served from magnetic-susceptibility measurements car-

Page 83: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

83A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

ried out during the warmup after demagnetization from2.5 T and 0.28 mK to zero field. Spin-glass-like phenom-ena were also detected.

Owing to its small Korringa constant, k590 msK, ex-change interactions are expected to be strong in scan-dium. Since the crystal structure is hcp and I5 7

2 for theonly stable isotope 45Sc, there is a quadrupole interac-tion between the crystalline electric-field gradient eqand the nuclear quadrupole moment Q . One then mustadd to the Hamiltonian of Eq. (1) the term

HQ5(i

@3e2qQ/4Ii~2Ii21 !#@Iiz2 2 1

3 Ii~Ii11 !# .

(59)

Pollack et al. (1992) have found that in zero magneticfield the ground state of Sc is 6 7

2, with the first excitedstate 6 5

2 located at 18 mK. Spontaneous nuclear order-ing of scandium can be described using an Ising system ifthe critical temperature is clearly below 18 mK.

Starting from the value of the nuclear magnetic mo-ment, m54.76mN , the dipole-dipole interaction HD canbe calculated; the result (Koike et al., 1995) predicts anantiferromagnetic transition at 130 nK. Estimates show,however, that the Ruderman-Kittel interaction HRK willproduce nuclear ordering at a higher temperature, in thelow microkelvin range.

A two-stage copper/scandium nuclear-demagnet-ization cryostat in Tokyo was used for these experi-ments. The sample was a 253333 mm3 single crystal,with the long edge parallel to the direction of the hex-agonal c axis. The second-stage demagnetizing field wasin the same direction. To produce zero magnetic field, amu-metal shield surrounded the sample. The magneticimpurities of the specimen were 3 ppm of Fe, 0.23 ppmof Cr, and 3.2 ppm of Mn. Although this impurity levelis not high, the residual resistivity ratio was only r(300K)/r(4.2 K) = 37, while the best value reported so far forscandium is 400. This reflects the large exchange en-hancement of Pauli paramagnetism and the strong ten-dency towards ferromagnetism in this metal.

The lowest temperature reached in the first nuclearstage was about 100 mK, attained by demagnetizingfrom 7 T. The ac magnetic susceptibility x was recordedusing a SQUID-based inductance bridge. The tempera-ture of the copper nuclear stage was measured by a Pt-NMR thermometer. A serious weakness of these experi-ments is that the temperature of the Sc nuclei wasinferred from indirect evidence only and never mea-sured.

Suzuki et al. (1994) first determined the temperaturedependence of x in zero magnetic field between 0.1 and10 mK. Below a dip at about 0.3 mK, the nuclear sus-ceptibility showed approximately Curie-like behavior.Later, the scandium single crystal specimen was cooledin a field of 2.4 T, keeping the first nuclear stage at 0.28mK for over 10 days, which should have been longenough to refrigerate the Sc spins to the same tempera-ture. Under these conditions, 90% of the entropyR ln2 of the ground-state doublet 6 7

2 was probably re-moved. Demagnetization from 2.4 T to zero field was

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

then carried out in 30 min. The time variation of themagnetic susceptibility parallel to the c axis of the Sccrystal during subsequent warmup was measured next.Two maxima in the x vs time curve were observed witha sharp dip at a higher temperature. Several demagneti-zation experiments from different starting temperatureswere performed with almost identical results; data foranother run, starting at 0.59 mK, are plotted in Fig. 99.The temperature, although not measured, is a monotoni-cally increasing function of time.

From the Curie-law behavior of x at higher tempera-tures, it was estimated that the upper peak in Fig. 99occurred around 70 mK. Using a polycrystalline sampleof 50 ppm Fe impurity, observed a shift of the maximumto 0.8 mK. They suggest that the dependence of thepeak position on impurity concentration shows that thehigher-temperature maximum of x is due to freezing ofthe electronic spin glass made of Fe impurities. It mightalso be possible that the lower-temperature peak is dueto spin-glass phenomena as well, either directly or indi-rectly. If this were the case, however, the peak shouldincrease proportionally to the Fe impurity concentra-tion, which was not observed. According to the authors,the most likely explanation is that the lower maximumin the magnetic susceptibility of scandium correspondsto nuclear ordering and to the formation of ferromag-netic domains.

Unfortunately, there are at least two serious difficul-ties in the interpretation of the data by Suzuki et al.(1994) and Koike et al. (1995): The temperature of thespecimen was not measured directly and the effect ofmagnetic impurities remains somewhat unknown.Therefore several uncertain conclusions had to be made.Further work on scandium is clearly warranted. To in-crease the reliability of the data interpretation, we feelthat it is also important to study how a quadrupolar sys-tem, with no cooperative ordering due to spin-spin inter-actions, behaves during demagnetization and the subse-quent warmup in a low field. Nevertheless, it is possible

FIG. 99. Time dependence of the magnetic susceptibility x ofscandium metal during warmup, after demagnetization toB50 from a 2.4 T external field parallel to the c axis of thehcp single crystal. The starting temperature before demagneti-zation was 0.59 mK. From Suzuki, Koike, Karaki, Kubota, andIshimoto (unpublished).

Page 84: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

84 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

that a ferromagnetic transition was reached in scandium.

C. Nuclear ordering in AuIn2

The ultralow-temperature group at Bayreuth hasmeasured the nuclear specific heat C , the nuclear mag-netic susceptibility x , and nuclear-magnetic-resonancespectra of AuIn 2 down to 30 mK. The large nuclearmagnetic moment (m55.5mN) and the high nuclear spin(I5 9

2), as well as the small Korringa constant (k590msK), would make indium metal, in principle, a favor-able candidate for studies of nuclear magnetic orderingphenomena. The low value of k ensures that there isthermal equilibrium in the sample at all times, i.e.,T5Te . However, the strong nuclear electric quadrupoleinteraction in tetragonal indium, producing a relativelylarge splitting (0.3 mK) of the nuclear hyperfine levelseven in zero external magnetic field (Symko, 1969;Karaki et al., 1994), and the rather high superconductingcritical field of 28 mT, preventing demagnetization tolower fields because thermal contact to electrons is lostin a superconductor, are severe drawbacks for studies ofnuclear ordering in indium metal itself.

To avoid these problems, the Bayreuth group has in-vestigated ordering of 115In nuclei in the cubic interme-tallic compound AuIn 2 (Herrmannsdorfer and Pobell,1995; Herrmannsdorfer, Smeibidl, Schroder-Smeibidl,and Pobell, 1995), which has a superconducting criticalfield of only 1.45 mT. Because of the small nuclear mag-netic moment of 197Au (m50.14mN), the contribution ofgold to the nuclear magnetic interactions in AuIn 2 isnegligible. In addition, the Korringa constant k5110msK of AuIn 2 is almost as small as for indium metal.Measurements of the heat capacity, magnetic suscepti-bility, and NMR spectra were made at 30 mK<T<10mK and at 2 mT<B<115 mT. Earlier NMR resultshave been reported by Gloos et al. (1990).

The AuIn 2 samples were prepared by melting 5N-pure gold and 6N-pure indium in a graphite crucible.The specimens were annealed for 40 h at 420°C; thisheat treatment resulted in a residual resistivity ratioRRR=500. From measurements of the static magneticsusceptibility an upper limit of 0.5 ppm was deduced forthe concentration of electronic magnetic impurities.X-ray diffraction data showed that the investigatedAuIn 2 samples were single crystals.

The experiments were performed in a copper nuclearrefrigerator (Gloos et al., 1988, 1991; Pobell, 1992b) intowhich sample holders made of silver were installed forthe NMR specimen, a calorimeter, a susceptometer, aswell as for a pulsed Pt-wire NMR thermometer. Formeasurements of the NMR spectra, a 431230.7 mm 3

piece of AuIn 2 was used, soldered with In to the speci-men holder. The calorimeter for the nuclear heat-capacity measurements was thermally isolated from itssurroundings by a superconducting heat switch made ofaluminum; the AuIn 2 sample was a cylinder, 17 mmlong and 5 mm in diameter. A set of three coils andniobium shields for the Pt-NMR thermometer, theAuIn2 sample, and the superconducting heat switch sur-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

rounded the calorimeter. The setup also contained a pairof coils for mutual-inductance measurements of thenuclear susceptibilities x8 and x9 of the AuIn 2 specimen.

The copper nuclear refrigerator cooled the calorim-eter and the AuIn 2 sample to the starting conditionsB5115 mT and T585 mK, which correspond to anentropy reduction DS/Smax50.35 and polarizationp5M/Msat50.74. The heat switch was then put into itsoff position and the AuIn 2 sample was demagnetized tothe low measuring field. The tiny heat leak, between 2and 10 pW, to the coldest parts of the apparatus and thesmall Korringa constant of AuIn 2 allowed good thermalequilibrium between the nuclei and the conduction elec-trons; the maximum difference DT between the sampleand the thermometer was 12% at the lowest tempera-tures and decreased rapidly as T was increased.

In the millikelvin temperature region, the NMR spec-tra of 115In showed a resonance peak at the nuclear Lar-mor frequency. The effective spin-spin relaxation timet2* , deduced from the NMR linewidths and measuredbetween 7 and 115 mT, was constant down to about 2mK, with values up to 600 ms, but then decreased sub-stantially to a minimum of 15 ms at 50 mK (see Fig. 100);this strong temperature dependence of t2* was rathersurprising and the first hint of a magnetic transitionnearby. For the Korringa constant, a field- (28 mT<B<115 mT) and temperature- (75 mK<T<10 mK)independent value k=112615 msK was deduced.

The nuclear magnetization of 115In was found fromthe free induction decay of the NMR signal, extrapo-lated to the middle of the excitation pulse. The resultsobtained in an external field of 28 mT are shown in Fig.101. The magnetization increased with decreasing tem-perature according to the Curie-Weiss law, reached amaximum at about 55 mK, and then decreased by abouta factor of 2 while the temperature was lowered to 30mK. The Weiss temperature u'30 mK. A similar T de-pendence was seen in 65- and 94-mT fields.

FIG. 100. Temperature dependence of the effective nuclearspin-spin relaxation time t2* of 115In in AuIn 2 at B528 and 65mT. The rather low saturation value of 230 ms for T.1 mK isdue to field inhomogeneities in the experimental setup. Withimproved arrangements relaxation times up to 600 ms wereobserved. From Herrmannsdorfer, Smeibidl et al. (1995).

Page 85: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

85A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

The mutual-inductance measurements of the nuclearsusceptibility, performed in 2-, 5-, and 13-mT fields,showed the same behavior. Data at 2 mT are plotted inFig. 101. The susceptibility changes again followed theCurie-Weiss law, but with u'43 mK; the maximum ofx was reached at 40 mK, indicating a phase transitionnearby. According to both types of magnetic measure-ments, the Weiss temperature is thus positive, whichshows that the magnetic interactions between thenuclear spins of In in AuIn 2 are predominantly ferro-magnetic.

The Bayreuth data on the nuclear specific heat ofAuIn 2, in an external magnetic field of 115 mT and be-tween 0.08 and 8 mK, lie nicely on the curve calculatedfor a nuclear paramagnet with the properties of nonin-teracting In moments. There was a small enhancementat 0.11 mK, which was more pronounced and occurredat lower temperatures in fields of 70 and 47 mT. How-ever, in an external field of 23 mT (see Fig. 102) onecould see, at all temperatures, an enhancement of theexperimental data over the curve for noninteractingnuclear spins; the heat-capacity maximum was at 45mK. This behavior was more pronounced in a 13-mTfield and most convincingly demonstrated by the datameasured in the smallest external field of 2 mT. Theheat-capacity maximum was 58 J/Kmol, which is almostthree orders of magnitude higher than the noninteract-ing value. The enhancement of C in the paramagneticrange corresponds to an internal field b57.0 mT actingon the In nuclei in AuIn 2. The temperature of the maxi-mum, TC53563 mK, is in excellent agreement with thesusceptibility data.

The ratio TC /u'0.82 is in good accord with the value0.72 calculated for a Heisenberg ferromagnet with largespins in cubic surroundings (Herrmannsdorfer and Po-bell, 1995). Similarly, the specific-heat enhancement,DC'55 J/K(mol of AuIn2)=27 J/K(mol of In), agreeswell with the prediction of the Heisenberg model giving

FIG. 101. Nuclear magnetization and susceptibility of 115In inAuIn 2. Open circles are results from ac mutual-inductancemeasurements at 16 Hz in a 2 mT field. Filled triangles aredata points deduced from NMR spectra at 260 kHz and 28 mT.From Herrmannsdorfer, Smeibidl et al. (1995).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

DC520 J/Kmol. The relative reduction of entropyDS/Smax50.09 at TC and in B52 mT is smaller than thevalue of about 0.14 for a simple cubic Heisenberg ferro-magnet with I5 9

2.The steep maximum in the nuclear specific heat (see

Fig. 102) near TC and shifts on the order of several kHzin the position of the NMR resonance both suggest thatthe phase transition may be of first order. Dipolar forcesbetween the In nuclei in AuIn 2 are on the order of 1mK. Hence the interactions must be dominated by ex-change coupling, which is presumably of the Ruderman-Kittel type. Using the measured Weiss temperatureu'43 mK, which equals 1

3I(I11)( jJ ij /kB in the mean-field theory, Herrmannsdorfer, Smeibidl, et al. (1995)find the parameter R552 for 115In in AuIn 2. The largevalue of R indicates a much stronger exchange domi-nance in AuIn 2 than, for example, in silver.

The Bayreuth measurements of the nuclear specificheat, nuclear magnetic susceptibility, and NMR spectraof 115In in AuIn 2, in the range of 30 mK<T<10 mKand 2 mT <B<115 mT, show that a ferromagnetic first-order phase transition occurs at the surprisingly highCurie point TC=35 mK. The Weiss temperature u543mK, and the internal field in the paramagnetic stateb510 mT. Many features of the data can be understoodwithin the nearest-neighbor, ferromagnetic Heisenbergmodel for a simple cubic ferromagnet with a large spin(Herrmannsdorfer and Pobell, 1995).

XIII. HYPERFINE-ENHANCED NUCLEAR MAGNETISMIN PRASEODYMIUM COMPOUNDS

In singlet ground-state ions like Pr 31, with high VanVleck susceptibilities, large hyperfine fields can be in-

FIG. 102. Nuclear specific heat C of 115In in AuIn 2, measuredin 23 mT and 2 mT external magnetic fields. The solid lines arecalculated for noninteracting In nuclei. The dashed curves inthe upper and lower figures are for Beff5(B21b2)1/2527 mT,b514 mT and for Beff57.3 mT, b57 mT, respectively. Theinset shows the behavior of C near TC534 mK. Modified fromHerrmannsdorfer, Smeibidl et al. (1995).

Page 86: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

86 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

duced at the nucleus by moderate external fields. WhenBext50, these ions have a nonmagnetic electronic singletground state of their 4f shell, but an applied fieldchanges the wave function and induces an electronicmagnetic moment on the ground state. This moment, inturn, produces a much stronger hyperfine field at thenucleus, which adds to the applied field. Enhancementfactors a511K , where K is the Knight shift, around20–100 are not uncommon. The apparent magnetic mo-ment of the nucleus is then one to two orders of magni-tude smaller than the Bohr magneton but two to oneorders of magnitude larger than the nuclear magneton.

The prerequisites for a high value of a are a smallseparation between the singlet ground state and the firstexcited levels, which leads to a high Van Vleck suscep-tibility, and a large hyperfine coupling constant. Rare-earth ions with integral values of the angular momentumJ suit these requirements best. Among them Pr 31 andTm 31 are particularly good because they have the low-est spin angular momentum S51. A small value of S isfavorable, since this reduces exchange interactions be-tween ions and thus makes spontaneous electronic po-larization less likely.

Use of hyperfine-enhanced materials for nuclear cool-ing has been reviewed by Andres and Lounasmaa(1982). This paper is still generally current as far asnuclear refrigeration is concerned; recent work hasmostly concentrated on matters important to nuclear or-dering.

A. Nuclear refrigeration

The possibility of hyperfine-enhanced nuclear refrig-eration was theoretically suggested by Al’tshuler (1966)and first put into practice by Andres and Bucher (1968).Reviews have been written by Lounasmaa (1974), byAndres and Lounasmaa (1982), and by Pobell (1992b).Andres and Bucher (1972 and references therein) haveinvestigated many intermetallic compounds in search ofsinglet ground-state behavior.

More often than not, however, they observed mag-netic order at liquid-helium temperatures, indicatingthat exchange interactions between the electronic mag-netic moments of the ions are strong enough to polarizethe singlet ground states. Van Vleck paramagnetism wasfound in intermetallic compounds of the type RX ,where R5Pr or Tm and X5Cu, Sb, Bi, Se, or Te. Indi-cations of nuclear refrigeration were seen in all thesematerials, but irreversible generation of heat, when theexternal magnetic field was swept downwards, oftenovershadowed the nuclear cooling effect and thus madethe outcome disappointing. Promising results were ob-tained with PrPt 5, PrCu 6, PrTl 3 , and especially PrNi 5.In these materials the spin-lattice relaxation time t1 isshort, but bulk thermal equilibrium is reached onlyslowly owing to the poor heat conductivity of thesamples. Buchal, Fischer et al. (1978) have also obtainedgood results for PrS. They were able to reach 0.72 mKby demagnetizing a sample made of this compound.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

An obvious advantage of hyperfine-enhanced nuclearrefrigeration is the strong polarizing field acting on thenuclei, which is usually at least ten times higher than theapplied field. In polycrystalline PrNi 5, which is a hex-agonal compound, the average hyperfine-enhancementfactor is 12. Although the Pr nuclear-spin density inPrNi 5 is smaller than Cu spins in copper, the coolingentropy per unit volume is about ten times larger in thepraseodymium compound. The local field B loc518 mTentering Eq. (13) is quite high.11 These properties makeit possible to build rather small refrigerators for reach-ing temperatures of 0.4 mK. Disadvantages when usingPrNi 5, on the other hand, are the not-so-ready availabil-ity of the material, its poor thermal conductivity, and therather high limiting low temperature.

Many cryostats based on hyperfine-enhanced nuclearcooling have been built, especially during the earlyeighties. Successful machines have been described byMueller et al. (1980), by Andres, Hagn et al. (1975), andby Greywall (1985). As an example of a modern cry-ostat, we show in Fig. 103 the apparatus of Greywall.

Recently nuclear refrigeration using copper has, onceagain, become more popular. An important advantageof brute force nuclear cooling with copper is the readyavailability of this metal in high-purity ingots, which re-

11Kubota et al. (1980) have found that B loc is not constant butincreases with field from the value quoted at B50.

FIG. 103. Low-temperature parts of the PrNi 5 nuclear refrig-erator of Greywall (1985). The PrNi5 bundle consists of seven8-mm diameter hexagonal rods, 95 mm long.

Page 87: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

87A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

sults in excellent thermal conductivity at low tempera-tures. The easily variable cooling capacity, proportionalto B2, and the wide temperature range available giveconsiderable flexibility to nuclear refrigerators based oncopper.

B. Nuclear ordering

Different types of Van Vleck paramagnets can beclassified using the parameter h52a2J(Q)/D , where Dis the energy separation between the ground state andthe first excited state and J(Q) is the Fourier transform,at wave vector Q, of exchange interactions between theions. If h,1 and if there are no hyperfine interactions,no magnetic ordering takes place in singlet ground-statesystems at any temperature, i.e., there is no induced mo-ment (Jensen and Mackintosh, 1991).

With hyperfine coupling Hhf5AI•J between thenuclear spins and the 4f electrons, two important effectsoccur. First, external fields felt by the nuclei are en-hanced as discussed above. Second, there is a magnetictransition even if h,1. The nature of the transition isvery different if h is near the critical value, i.e., close toone, in comparison with the situation in which h!1.

For nearly critical h , the phase change can be de-scribed as a nuclear-induced electronic transition. Thetheory of this process has been developed by Murao(1972, 1981). The nuclei remain disordered just belowTc and align only at lower temperatures. Ordering tosuch a mixed electron-nuclear state has been observedin PrCu 2 below 50 mK (Andres, Bucher, Maita, andCooper, 1972) and in PrCu 5 below 40 mK (Andres,Bucher et al., 1975; Genicon, Tholence, and Tournier,1978). These compounds have also been investigated inneutron-diffraction experiments by Benoit et al. (1981)and Nicklow et al. (1985). Pure praseodymium metal ap-pears to be a mixed electron-nuclear system. Early stud-ies of the magnetic behavior of praseodymium producedcontroversial results, presumably due to magnetic impu-rities in the specimens. Cooperative magnetic orderingat Tc525 –30 mK was established in the heat-capacitymeasurements by Lindelof, Miller, and Pickett (1975). Asinusoidal modulation of the nuclear magnetic momentin Pr has been observed in neutron-diffraction measure-ments below 60 mK (Kawarazaki et al., 1988; McEwenand Stirling, 1989).

For h!1, the hyperfine-enhanced system can best bedescribed as a nuclear magnet in which the moments,enhanced by the factor 11K , are coupled by an indirectexchange interaction. Magnetism in such a spin assem-bly is similar to that found, for example, in copper atnanokelvin temperatures. The exchange interaction inenhanced nuclear magnets can be thought of as follows:The magnetic nucleus in a Pr ion i induces a 4f elec-tronic moment in that ion. This moment then couples tothe 4f electrons in a neighboring ion j by the Ruderman-Kittel-Kasuya-Yosida interaction and, in turn, tonucleus j through the hyperfine interaction (Rudermanand Kittel, 1954; Kasuya, 1956; Yosida, 1957).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

In summary, the energy scales for Pr 31 ions are suchthat the ordering is purely of nuclear origin only if thetransition temperature is below 10 mK. For Tc’s in therange between 10 mK and 1 K, the ordering is to amixed electron-nuclear state. Finally, if Tc is above 1 K,the transition is probably purely electronic.

Among praseodymium compounds for which h!1,ferromagnetic ordering has been observed at TC52.5mK for PrCu 6 (Babcock et al., 1979) and at TC50.40mK for PrNi 5 (Kubota et al., 1980). Nonferromagnetic,presumably antiferromagnetic nuclear ordering has beenseen at TN50.5 mK in PrSe (Kubota et al., 1987). Thesethree compounds are the only metals, besides dilutedPrNi 5, for which hyperfine-enhanced spontaneousnuclear magnetic ordering has been reported.

The other compounds in the series, PrS, PrSe, andPrTe, are also interesting systems for studies of nuclearmagnetism. Kubota et al. (1984) measured the ac suscep-tibility of PrS down to 80 mK, but they observed noevidence for magnetic ordering. PrS, PrSe, and PrTehave the simple NaCl structure in which Pr nuclei oc-cupy fcc sites. It would be interesting to study the spin-spin interactions and ordered magnetic structures ofthese metals in more detail, as well as to compare themwith the properties of Cu, Ag, and Rh.

The specific-heat data for PrCu 6 and PrNi 5 are illus-trated in Fig. 104. A sharp anomaly is observed at TC ,which coincides with a peak in the magnetic susceptibil-ity vs temperature curve. Figure 104 also shows resultsfor PrCu 2 and PrCu 5, in which the transition takes placein a mixed nuclear-electronic state. The different char-acter of the ordering is evident: The specific-heat peakin PrCu 2 and PrCu 5 is broad and the magnetic suscep-tibility reaches its maximum value at a higher tempera-ture than does the specific heat.

Other investigations of PrNi 5 by the Julich group in-clude the finding that the ratio of the Curie constant tothe saturation magnetization is larger than what onewould expect for localized hyperfine-enhanced moments

FIG. 104. Reduced specific heats C/R of PrNi5 , PrCu6 ,PrCu5 , and PrCu2 vs temperature. The arrows indicate thepositions of the maxima in the susceptibility vs temperaturecurves for the corresponding compounds. The figure has beenmodified from that of Babcock et al. (1979) by including thedata for PrNi5 from Kubota et al. (1980).

Page 88: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

88 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

on the basis of the known enhancement factor11K512 (Kubota et al., 1983). This was taken as evi-dence for extra fluctuating magnetic moments that havea long wavelength and do not give large contributions tothe specific heat at low temperatures.

The group at Osaka has measured the electrical resis-tivity r of PrCu 6 across the magnetic phase transition(Miki et al., 1992). A decrease in r started at TC=2.6mK, which was attributed to ferromagnetic ordering.Resistivity also displayed critical phenomena just aboveTC .

The influence of magnetic dilution on the electronicand nuclear magnetic properties of Pr 12xY xNi 5 was in-vestigated recently by Herrmannsdorfer, Uniewski, andPobell (1994a, 1994b) for yttrium concentrations x50,0.02, 0.05, 0.10, 0.20, and 1.00. Replacement of Pr 31 byY 31 was expected to have a substantial effect on theelectronic and nuclear magnetic properties of these com-pounds because yttrium lacks the 4f electron. Themutual-inductance measurements of the 16-Hz nuclearsusceptibilities were made in the temperature range 50mK<T<8 mK and in fields Bext<0.1 mT. Each of thesix specimens had a mass of about 0.1 g.

The nuclear susceptibility x of Pr 12xY xNi 5 is plottedas a function of temperature in Fig. 105 between 50mK and 1.0 mK for three yttrium concentrations. In theparamagnetic region at high temperatures, the 1/x vs Tplots display a typical ferromagnetic behavior. AtT,0.6 mK, a substantial increase of x is observed foreach compound, indicating a spontaneous nuclear mag-netic ordering transition. The maximum of each curvewas assumed to define the Curie temperature TC . Allsamples containing praseodymium showed a nuclearphase transition, with TC reduced from 370 mK at x50to 100 mK at x50.20. The substantial depression of theordering temperature might make the diluted PrNi 5compounds useful for nuclear refrigeration into the mi-crokelvin region.

FIG. 105. Nuclear magnetic susceptibility of Pr12xYxNi5 below1.0 mK as a function of temperature: s , x50; n , x50.05;h , x50.20. The maxima of the curves occur approximately atthe nuclear Curie temperature TC . From Herrmannsdorfer,Uniewski, and Pobell (1994a, 1994b).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

The ratio of the observed nuclear Curie and Weisstemperatures depends strongly on the yttrium concen-tration: TC /u increases from 1.7 for x50 to 7.7 forx50.20. The measured values of TC(x) are in reason-able agreement with those calculated from the mean-field theory.

Moyland et al. (1995) have recently measured thezero-field nuclear ac susceptibility of PrBe 13 from 3 mKto 20 mK. Their data indicate antiferromagnetic Curie-Weiss behavior with the Weiss temperature u522.2mK.

Various theoretical aspects of hyperfine-enhancednuclear magnets have been investigated extensively byIshii and co-workers (see, for example, Ishii andAoyama, 1991; Akai and Ishii, 1994).

XIV. SPONTANEOUS NUCLEAR ORDER IN INSULATORSAT T>0 AND T<0

Owing to the poor thermal conductivity of insulatorsand their long intrinsic spin-lattice relaxation times,studies of nuclear ordering in dielectric materials arequite different from corresponding investigations onmetals. Abragam and Goldman and their co-workers atSaclay (Chapellier, Goldman, Chau, and Abragam,1970; Abragam and Goldman, 1982; Abragam, 1987)have conducted experiments that clearly demonstratenuclear antiferromagnetism in CaF 2 and nuclear domainferromagnetism in LiH, both materials being insulatorsof cubic crystal structure with spin I5 1

2 of the 19F or1H nuclei. Wenckebach and co-workers (Marks, Wenck-ebach, and Poulis, 1979; Van der Zon, Van Velzen, andWenckebach, 1990) have investigated Ca(OH) 2 atT,0. Several aspects of these studies are worthy of ashort digression from our main topic. We also refer thereader to a recent review by Bouffard et al. (1994).

A. Dynamic nuclear polarization

First, it should be noted that for studies of nuclearcooperative phenomena in insulators it is neither neces-sary nor is it possible to refrigerate the lattice to a tem-perature of 1 mK or below. It is sufficient just to cool thenuclei; the very long spin-lattice relaxation time t1, es-pecially in insulators, again ensures that after demagne-tization the nuclei are effectively decoupled from therest of the specimen and thus remain cold long enoughfor experiments to be carried out.

The magnet used by Chapellier et al. (1970) generateda field B52.7 T. With the internal field b'0.2 mT at thesite of the 19F nuclei in CaF 2, it was found thatTf /Ti5731025 if demagnetization were carried outfrom Bi52.7 T all the way to Bf50. A starting tempera-ture Ti,10 mK would thus be necessary for reachingthe microkelvin range. Although this temperature caneasily be produced today by means of a suitable precool-ing stage, it is doubtful whether the fluorine nuclei couldever be cooled in equilibrium with the lattice, even withthe help of electronic paramagnetic impurities, to the

Page 89: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

89A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

vicinity of 10 mK because of the extremely long relax-ation time. The problem had to be solved in a differentway.

The CaF 2 specimen, a sphere of about 1.5 mm in di-ameter, was first cooled at the tip of a copper cold fingerto 0.7 K by means of a 3He refrigerator. The 19F nucleiwere then dynamically polarized by a method known asthe ‘‘solid effect,’’ which now will be explained with thehelp of Fig. 106.

Let us consider a specimen in which a few electronicmoments are mixed with the magnetic nuclei; this wasachieved in the experiments of Chapellier et al. (1970)by introducing U 31 (or Tm 21) ions, at a concentrationof 10 24, as paramagnetic impurities into the CaF 2 lat-tice. At T50.7 K and Bi52.7 T, the polarization of thenuclei was almost zero, whereas the three orders of mag-nitude larger electronic magnetic moments were nearlyfully polarized (p'1). Next, the system was pumped bya microwave field of angular frequency ve2vn , equal tothe difference between the Larmor frequencies of theelectronic and nuclear spins. Such a field can induce‘‘flip-flop’’ transitions, the required energy being ab-sorbed from the field of microwave power. Starting fromthe situation depicted in Fig. 106(a), the first time theelectronic spin flips, one of the nuclear spins that origi-nally pointed down will reverse its direction as in Fig.106(b); an energy quantum \(ve2vn) is thereby ab-sorbed by the spin system.

Owing to its short relaxation time with the lattice, theelectronic spin quickly returns to its original direction,by simultaneously transferring an energy quantum \veto the lattice, while the nuclear spin, because of its verylong relaxation time, will remain in its new direction asin Fig. 106(c). The process is then repeated, and, one byone, the nuclei in the vicinity of the U 31 impurity be-come polarized in the direction of the external magneticfield. By spin diffusion the polarizing effect of each elec-tronic impurity gradually spreads out.

An interesting feature of the solid effect is that it canbe used just as well for producing negative absolute tem-peratures in the nuclear-spin system. If the microwavefrequency employed for pumping equals ve1vn , a pairof electronic and nuclear moments pointing originally inthe same direction will reverse together in a flip-fliptransition. The end result [see Figs. 106(d)–106(f)] isthat the nuclei become polarized in the direction oppo-

FIG. 106. Positive (upper sequence) and negative (lower se-quence) nuclear-spin temperatures produced by the ‘‘solid ef-fect.’’ Long arrows depict electronic and short arrows nuclearmagnetic moments. For further details, see text.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

site to the external magnetic field Bi . There are thusmore nuclei in the higher energy level than in the lowerone, i.e., the spin temperature T is negative (see Fig. 8).

In practice, the Saclay group produced up to 90%nuclear polarization in CaF 2 after three hours of micro-wave pumping at 0.7 K. The power was then turned off,which allowed the system to cool to 0.3 K, the tempera-ture of the 3He precooling stage. At 0.3 K the nuclearpolarization p decayed sufficiently slowly for experi-ments. It should be noted that p50.5 in a 2.7-T magneticfield corresponds to a spin temperature T564 mK [seeEq. (25a)], the sign depending on the direction of polar-ization.

B. Adiabatic demagnetization in the rotating frame

Having thus reached the desired starting conditionsfor nuclear cooling, Ti564 mK and Bi52.7 T, the nextstep in the experiment of Chapellier et al. (1970) was todemagnetize. One could, of course, reduce the externalfield to zero in the hope of reaching very low positive ornegative temperatures in the nuclear-spin system, de-pending on whether Ti.0 or ,0, respectively. Thissimple procedure, however, is unsatisfactory because ininsulators with paramagnetic ions the spin-lattice relax-ation proceeds through impurities. In a low externalfield, the energy difference \(ve2vn) becomes small inabsolute value and direct energy exchange between thenuclear- and electronic-spin systems can occur easily.The cold nuclei would thus reach equilibrium rather rap-idly with the far hotter electronic impurities and losetheir order quickly. To avoid these difficulties, demagne-tization was performed in a rotating frame of reference,which we now describe.

It was already mentioned that after the microwavepumping was stopped the sample cooled to 0.3 K in aconstant external field Bi52.7 T. One assumes that thefield is pointing in the z direction. In this field the 19Fnuclei have a Larmor frequency vn /2p5107 MHz. Onethen applies to the specimen a small external magneticfield, b0'5 mT, rotating in the xy plane, i.e., at rightangles to Bi , with an angular frequency v0 that is of thesame order of magnitude as vn . In the frame rotatingwith b0, the Larmor frequency of the nuclear spins ap-pears to be (vn2v0)/2p , i.e., the magnetic field that thenuclei see in the z direction is (vn2v0)Bi /vn . The signof the initial temperature Ti is then positive or negativedepending on whether the initial magnetization is paral-lel or antiparallel to Bi .

In the rotating coordinate frame, b0 appears tobe constant in its direction and magnitude and at rightangles to Bi . The nuclear spins thus experience aneffective field whose magnitude is Beff5$@(vn2v0)/gn]21b0

2%1/2. Adiabatic demagnetization can nowbe performed by sweeping the field (or the frequency)from Bi5vn /gn to Bf5v0 /gn ; here gn is the gyromag-netic ratio. The minimum value of the effective magneticfield, Beff5b0, is reached at resonance vn5v0. If dipole-dipole and exchange interactions are ignored the

Page 90: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

90 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

nuclear-spin system should now reach a temperatureTf5(b0 /Bi)Ti'610 nK, the sign of Tf being the sameas the sign of Ti564 mK. The actual final temperatureis, however, between 0.1 and 1 mK (or between –0.1 and–1 mK) owing to dipolar interactions between nuclearspins. The final demagnetization of b0 is then per-formed, followed by remagnetization in the sequence 5mT→0→5 mT; this is the well-known adiabatic fast-passage technique.

Similarly to the situation in the laboratory frame, thefinal field after demagnetization in the rotating coordi-nate system can be nonzero. This is the case if vn is notswept all the way to vn5v0. Nuclear spins will thenexperience an effective external field that competes withthe dipolar forces.

For the U 31 electronic moments at Bi52.7 T, theLarmor frequency ve/2p570 GHz@vn/2p . In the framerotating at the nuclear Larmor frequency v0/2p , the ef-fective field on electronic moments hardly changes at all.Demagnetization in the rotating frame is thus ‘‘selec-tive:’’ electronic moments stay hot while nuclear spinsare cooled. Nuclei maintain their high degree of orderfor a long time, since direct energy exchange betweenthe nuclei and the electrons does not occur because\(ve2vn) is large.

C. Truncated dipolar Hamiltonian

In the rotating coordinate frame, spins feel only thetruncated part of the dipolar interaction, viz.,

Htrunc5(i,j

aij* ~2IizIj

z2IixIj

x2IiyIj

y!, (60)

aij* 5 12 \2g ig jr ij

23@123~cosu ij!2# , (61)

where u ij is the angle between rij and Bi ; here rij is thelattice vector from site i to site j . Thus the orientation ofthe external field determines the spin-spin coupling aswell as the ordered ground state.

This can be predicted using the method described inSec. XV.B. One important characteristic of the problemis the distinction between longitudinal and transversespin configurations with respect to the external field.Both types of structures have been observed, as will bediscussed below. Many features of longitudinal struc-tures could be described in terms of an Ising system withlong-range interactions. In the transverse configurationsonly the components normal to Bi are ordered. There-fore, although the underlying crystal structure is cubic,magnetic ordering in the rotating frame never displayfeatures associated with cubic symmetry, which are im-portant for spin structures in copper and silver.

D. Experimental results on CaF2, LiH, and Ca(OH)2

The first and most extensively studied dielectric com-pound for nuclear ordering was CaF 2 (Chapellier, Gold-man, Chau, and Abragam, 1970). The nuclear magneticmoment of 19F is almost as large as that of 1H, whereasmost calcium nuclei are of the spinless isotope 40Ca. The

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

behavior of the 19F nuclei after demagnetization was in-vestigated using the adiabatic fast-passage technique,i.e., by recording the NMR dispersion curve whilesweeping the field b055 mT at the rate db0 /dt510mT/s. At b050 the height of the dispersion signal isproportional to the transverse susceptibility x' . TheNMR absorption signal x9 was employed for determin-ing the nuclear polarization before and after each fastpassage.

Figure 107 shows experimental results by the Saclaygroup on the transverse susceptibility with Bi parallel toa [100] direction of the CaF 2 single crystal. At T,0 andat initial polarizations upu.0.3, x' becomes constant, in-dicating that the nuclei order antiferromagnetically. Itwas not possible to determine the transition tempera-ture. At T.0, the susceptibility behaves, up to p50.45, in a way that is characteristic of paramagnetism.An antiferromagnetic structure is expected in this caseas well, but with a higher critical polarization.

A variety of other kinds of spin structures were ob-served in CaF 2 depending on the sign of T and the di-rection of the external magnetic field. A great deal ofinformation was obtained by using the magnetic isotope43Ca, with the concentration of 0.13% in natural cal-cium, as a probe. For the [111] alignment of the externalmagnetic field, a domain ferromagnet was observed atT,0. At T.0, the spin arrangement was found to be ahelix in which spins precess in the plane perpendicularto Bi at the Larmor frequency while preserving theirrelative orientations (Urbina et al., 1982, 1986).

The Saclay group also investigated nuclear magneticordering in single crystals of LiH by means of neutron-diffraction techniques (Roinel et al., 1978, 1980, 1987).This compound is a suitable choice because of its simplecubic structure and because the spin-dependent scatter-ing cross section is large for protons. Figure 108 illus-

FIG. 107. Transverse susceptibility of the 19F spin system inCaF 2, measured in the rotating frame, as a function of polar-ization: s , at positive nuclear temperatures; m, at negativenuclear temperatures. From Chapellier et al. (1970).

Page 91: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

91A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

trates the antiferromagnetic spin configurations atT.0 and at T,0, with Bi parallel to the [001] crystal-line axis, deduced from the neutron data. Both struc-tures are longitudinal since spins are either parallel orantiparallel to Bi . For T,0 and with Bii@110# , thestructure is a domain ferromagnet.

The Leiden group of Wenckebach has investigatednuclear magnetic ordering of protons in Ca(OH) 2. Theexperimental method was the same as that employed atSaclay. Data were obtained at negative spin tempera-tures with the external magnetic field parallel to thecrystalline c axis (Marks et al., 1979; Van der Zon et al.,1990). The results, illustrated in Fig. 109, show that, forupiu,0.3, the nuclear-spin structure is paramagnetic,while at larger polarizations x' is nearly independent ofpi , agreeing approximately with the curve calculated forthe longitudinal domain structure with ferromagnetic or-der. TC520.960.2 mK was deduced for the Curie tem-perature.

The Leiden group (Van Kesteren et al., 1985) also in-vestigated the possibility of removing two of the greatestdrawbacks of the dynamic nuclear polarization method:

FIG. 108. Nuclear antiferromagnetic structures in LiH whenthe external field is aligned along a [100] axis: black arrows7Li; white arrows 1H. From Roinel et al. (1978).

FIG. 109. Transverse susceptibility x' vs the initial polariza-tion pi before adiabatic demagnetization in the rotating frame.The curves were calculated using the restricted-trace approxi-mation. Modified from Van der Zon, Van Velzen, and Wenck-ebach (1990).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

(i) the relatively rapid destruction of the nuclear order,through spin-lattice relaxation, by the large local fieldsproduced by the needed electronic impurities, U 31 orTm 21 ions; and (ii) the influence of these local fields onthe long-range nuclear-spin configurations. It was hopedthat the use of microwave-induced optical nuclear polar-ization (MIONP) (Deimling et al., 1980) would removeboth problems. A suitable substance for trying the MI-ONP method is fluorene, C 13H 10 , doped withphenantharene C14D10. UV light photo-excites thephenantharene molecules to their lowest triplet state,creating electronic spins S51. After the 1H nuclei arepolarized by the solid effect, the light is turned off andthe phenantharene molecules decay to their ‘‘harmless’’diamagnetic ground state. With this method, Van Kes-teren et al. (1985) achieved proton polarizations p50.42. Unfortunately, it turned out that this was notenough to produce nuclear ordering, and further devel-opment of the MIONP technique has been abandoned.

XV. THEORY

In this long section we first describe (Sec. XV.A) thetheory of spin-spin interactions mediated by conductionelectrons. In Sec. XV.B we present the mean-field de-scription of magnetic ordering which provides the basicframework for the entire chapter. Section XV.C com-pares measured magnetic properties, such as the criticaltemperature and the behavior of susceptibility and en-tropy with temperature, of Cu, Ag, and Rh nuclei withtheoretical calculations.

An exhaustive discussion of ordered spin structuresthen follows. We first deal with type-I antiferromag-netism in Sec. XV.D. Since the mean-field theory doesnot yield a unique ground state, one has to consider ef-fects due to thermal and quantum fluctuations. Theground-state spin configuration depends sensitively onthe direction of the external magnetic field because ofthe anisotropic dipolar interaction. The spin structuresare first calculated for fields along the crystalline high-symmetry directions. The results are then comparedwith neutron-diffraction measurements on copper,where type-I order has been found in the high-field re-gion.

In Sec. XV.F we thoroughly discuss the (0 23

23) spin

structures observed for copper in intermediate externalfields. A relatively simple mean-field description repro-duces the neutron-diffraction results. The problem ofthe so-far-unknown ordering of copper nuclei in highfields aligned close to the [111] crystalline direction isdiscussed in Sec. XV.G. Finally, in Sec. XV.H, we re-view the theory of the ferromagnetic domain structureobserved in silver at negative temperatures.

A. Exchange interactions

Indirect exhange interactions mediated by conductionelectrons are crucial in determining the structure of or-dered spin configurations in metals. The force between

Page 92: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

92 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

the nuclear spins arises from the hyperfine interactionbetween an electron and a nuclear spin, as was firstshown by Ruderman and Kittel (1954) and soon after-wards by Bloembergen and Rowland (1955). The inter-action can be understood in terms of two processes: Onenuclear spin scatters a conduction electron into an ex-cited energy level, and another spin then scatters theelectron back to its initial state. This virtual processleads to an effective coupling between the two nuclearspins.

The theory of indirect exchange was quickly appliedto magnetic coupling between ion cores by Kasuya(1956) and by Yosida (1957). Such an interaction is con-sidered particularly important in the rare-earth metals.We refer the reader to the review by Kittel (1968) forthe vast number of publications written on this topic.

1. Ruderman-Kittel interaction

Let us first recall the original result of Ruderman andKittel (1954), which was derived using the free-electronapproximation. By applying second-order perturbationtheory, they showed that the contact interaction

Hel-ncont5

8p

3\2gngeI•Sd~r! (62)

between the nuclear spin I and the electron spin S, withthe respective gyromagnetic ratios gn and ge , leads tothe isotropic spin-spin interaction

HRK52 12 (

i ,jJ ijIi•Ij, (63)

where

Jij5hm0

4p

\2gn2

rij3 Fcos~2kFrij!2

sin~2kFrij!

2kFrijG . (64)

The coupling constant Jij oscillates with distance and hasa long range. The coefficient12 h is negative and dependson the density of conduction electrons at the site of thenucleus as uc(0)u4. The free-electron model is able toyield only slightly better than an order-of-magnitude es-timate for h . Much more accurate predictions can beobtained from first-principles electronic band-structurecalculations, at least in simple elemental metals such ascopper and silver.

The first state-of-the-art band-structure calculation ofthe Ruderman-Kittel (RK) interaction in copper wasmade by Lindgard, Wang, and Harmon (1986). Ratherthan calculating Jij in real space, these authors first com-puted the Fourier transform

J~q!564p2

9\4gn

2ge2 1

N (n ,n8,k

fn~k!@12fn8~k1q!#

En8~k1q!2En~k!

3ucn ,k~0 !u2ucn8,k1q~0 !u21constant , (65)

12The relationship between h and R of Eq. (6) isR50.587h .

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

where fn(k) is the Fermi function corresponding to theband energy En(k) for a wave vector k in the first Bril-louin zone with a band index n , and ucn ,k(0)u2 is theelectron density at the nucleus. Only s electrons with aspherical charge distribution have a nonvanishingucn ,k(0)u2. The constant is the ‘‘self-energy’’ term, whichensures that Jij50 for rij50 when J(q) is Fourier trans-formed back to real space, i.e., (qJ(q)50.

Lindgard et al. (1986) found that the most significantcontributions to J(q) arise from the pairs of bands(n ,n8)5(1,6) and (1,7), which have strong s-like char-acter, while the contribution from (n ,n8)5(6,6) nearthe Fermi level was less significant. The authors notedthat this is contrary to the free-electron calculation ofRuderman and Kittel (1954) in which the charge densi-ties in Eq. (65) were assumed constant and equal to anaverage value at the Fermi level.

Lindgard et al. calculated J(q) for 14 different valuesof q. The results were then fitted to an eight-nearest-neighbor model to obtain the interaction in real space.The calculated Jij’s are given in Table IV. A comparisonwith the free-electron model shows that the nearest-neighbor coupling clearly dominates the Jij values calcu-lated by the band theory. The Jij’s also show an oscilla-tion with rij but with a much reduced amplitude.

The RK interaction in copper has also been computedfrom the electronic band structure by Frisken and Miller(1986, 1988b). Their method of calculation yielded theinteraction directly in real space and emphasized preciseintegration of the energy denominator in Eq. (65). Theaccuracy of the matrix elements, ucn ,k(0)u2, in their firstpaper (1986) was, however, criticized by Harmon andWang (1987). Frisken and Miller later improved thetreatment of the matrix elements (Miller and Frisken,1988). Results of their nonrelativistic orthogonalizedplane-wave calculations are included in Table IV. Thedata agree reasonably well with those of Lindgard,Wang, and Harmon (1986), while the discrepancy withthe traditional RK result is large.

It is perhaps surprising that relativistic effects are sig-nificant for the RK interaction even in an element aslight as copper. Corrections to the band structure aresmall but they have a significant influence on the radialfunctions of s electrons near the nucleus. A self-consistent, scalar relativistic calculation for atomic cop-per, using the proper relativistic hyperfine operator,shows that atomic hyperfine splitting is increased by14% in comparison with the corresponding nonrelativis-tic value (Oja, Wang, and Harmon, 1989). The RK in-teraction in bulk copper is enhanced by approximatelythe same amount by relativistic corrections, as is shownin Table IV.

The RK interaction in silver has been calculated usingthe same techniques as in copper (Miller and Frisken,1988; Harmon et al., 1992). The results are presented inTable V. The general features are the same as in thecase of copper. The interaction is more clearly domi-nated by the nearest-neighbor coupling than the free-electron result, although the sign of Jij oscillates simi-

Page 93: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

93A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Rev. Mod. Phys

TABLE IV. Ruderman-Kittel coupling strength Jij for copper according to various calculations. Thevalues are listed in nK.

rij LWHa FMb OWHc Free el.d

(1,1,0) 211.67 28.99 212.74 211.60(2,0,0) 1.41 1.56 1.63 5.18(2,1,1) 21.31 20.76 21.88 22.88(2,2,0) 0.11 20.67 20.17 20.36(3,1,0) 0.25 0.61 0.56 1.47(2,2,2) 0.61 0.34 0.94 0.25(3,2,1) 20.06 20.13 0.08 20.83(4,0,0) 20.18 20.17 20.35 20.49(4,1,1) 20.03 20.27

R 20.34 20.25 20.37Rexp520.4260.05e

Q 0.091 0.070 0.101Qexp50.09560.003f

aNonrelativistic calculation of Lindgard, Wang, and Harmon (1986).bNonrelativistic calculation of Frisken and Miller (1988b).cRelativistic calculation of Oja, Wang, and Harmon (1989).dFree-electron approximation with overall scaling to yield R520.42.eFrom NMR measurements on highly polarized spins by Ekstrom et al. (1979).fNMR linewidth measurement of Andrew et al. (1971).

larly in both cases. The relativistic corrections enhancethe RK coupling in silver by 40%.

The theoretical values for Jij can be directly comparedwith two experimentally determined parameters. Thestrength of the RK interaction is often described by thedimensionless quantity

TABLE V. Ruderman-Kittel coupling strength Jij for silveraccording to various calculations. The values are listed in nK.

rij MF a HWL b

(1,1,0) 20.895 21.335(2,0,0) 0.065 0.193(2,1,1) 20.145 20.193(2,2,0) 20.051 20.031(3,1,0) 0.086 0.077(2,2,2) 0.127 0.118(3,2,1) 20.013 20.009(4,0,0) 20.041 20.055(4,1,1) 20.008

R 21.53 22.26Rexp522.560.5c

Q 0.406 0.598Qexp50.55360.031d

aNonrelativistic calculation of Miller and Frisken (1988).bRelativistic calculation of Harmon, Wang, and Lindgard

(1992).cFrom NMR measurements on highly polarized spins (Oja,

Annila, and Takano, 1990; Hakonen, Yin, and Lounasmaa,1990; Hakonen, Nummila, and Vuorinen, 1992).

dNMR linewidth measurement of Poitrenaud and Winter(1964).

., Vol. 69, No. 1, January 1997

R5(j

J ij /~m0\2g2r!, (66)

which can be deduced from various NMR and suscepti-bility measurements on highly polarized spins (see Sec.XI). A positive R is characteristic of ferromagnetic in-teractions, whereas a negative R indicates antiferromag-netism.

Experiments in the limit of a small nuclear-spin polar-ization yield the parameter

Q5S (j

J ij2 D 1/2Y~m0\2g2r!, (67)

which describes the average strength of the local fluctu-ating field. In a system with two or more spin species,measurements of the second moment of the NMR ab-sorption line (Van Vleck, 1948) yield Q . Such an experi-ment is straightforward in silver (Ruderman and Kittel,1954; Poitrenaud and Winter, 1964) but difficult in cop-per because the second moment is mainly determined bythe dipolar interaction. Using the magic-angle spinningtechnique it is, however, possible to remove the dipolarcontribution from ^Dn2& and to obtain Q accurately(Andrew et al., 1971; Andrew, 1973; Andrew and Hin-shaw, 1973).

2. Anisotropic exchange interactions

In addition to the contact interaction, there are twoother contributions to the hyperfine forces (Lindgrenand Rosen, 1974a, 1974b), viz., the dipolar interactionbetween the nuclear and electronic moments

Hel-ndip 52\2gnger23@I•S23r22~I•r!~S•r!# , (68)

Page 94: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

94 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

and the orbital interaction

Hel-norb 5\2gnger23I•lW . (69)

Here r is the vector separating the nucleus and the elec-tron and lW is the electronic angular momentum. BothHel-n

dip and Hel-norb contribute to the indirect nuclear-spin

coupling, as was first shown by Bloembergen and Row-land (1955). The resulting force is, in general, aniso-tropic in spin space. In the so-called Bardeen’s (1937)spherical approximation for the electronic band struc-tures, which was employed by Bloembergen and Row-land, the resulting interaction can be divided into an iso-tropic term like the RK coupling and an anisotropicterm with the symmetry of the dipolar interaction. Theanisotropic term is therefore often known as the pseudo-dipolar interaction. The terminology is, however, mis-leading since the dipolarlike symmetry results only fromthe use of the spherical approximation for theband structure. It would be more appropriate tocall the forces arising from Hel-n

dip and Hel-norb non-s-

electron-mediated interactions, since s electrons do notcontribute to them.

The relative importance of the traditional RK termmediated by the contact interaction, and hence only bys electrons, obviously depends on the partial-wave char-acter of conduction electrons. In some transition metals,such as platinum and lead, anisotropic interactions notmediated by s electrons are comparable to isotropicforces (Froidevaux and Weger, 1964; Alloul and Froi-devaux, 1967). In copper and silver, which have one un-paired s electron and filled d shells in their atoms, onewould expect a dominating RK interaction.

Interestingly, however, the d-electron terms in coppercontribute as much as 20% to the measured spin-latticerelaxation rate through the orbital hyperfine interaction(Asada et al., 1981; Ebert et al., 1984). This led Oja andKumar (1987) to investigate the role of forces not medi-ated by s electrons. They concluded that orbital interac-tions are important in the nuclear-spin coupling to theextent of possibly changing the ground-state orderingvector from what it would be without these terms. How-ever, since the band structure was modeled only in termsof the crude spherical approximation (Bardeen, 1937), amore sophisticated study was needed to settle the ques-tion.

First-principles band-structure calculations for all in-teractions not mediated by s electrons were performedby Oja, Wang, and Harmon (1989). Using scalar relativ-istic wave functions obtained from a linear, augmented-plane-wave calculation, they found that orbital interac-tions do make a significant contribution to exchangeforces between copper nuclei, as was suggested by Ojaand Kumar (1987). The theoretical nearest-neighbor in-teraction matrix was found to be

Aex@rij5~a ,a ,0!#/kB5S 212.6 22.0 0

22.0 212.6 0

0 0 29.4D ,

(70)

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

where the units are nK. Interactions not mediated by selectrons are such that they decrease the theoretical val-ues for R and Q by 10%. Deviations from the fully iso-tropic interaction are on the order of 20% (see TableIV). The symmetry of calculated anisotropic interactionsis not dipolar in form but is determined, in general, onlyby the symmetry of the lattice (Griffiths et al., 1959; Ojaet al., 1989).

The anisotropy of exchange interactions in copper is,however, small in comparison with the anisotropy of thedipolar interaction. For nearest neighbors

Adip@rij5~a ,a ,0!#/kB5S 12.7 38.1 0

38.1 12.7 0

0 0 225.4D . (71)

The full nearest-neighbor interaction, Aij5Aijex

1Aijdip , is

A@rij5~a ,a ,0!#/kB'S 0 36 0

36 0 0

0 0 235D , (72)

emphasizing the fact that the mutual spin forces in cop-per are, indeed, very anisotropic.

Interactions not mediated by s electrons have notbeen computed for silver. The calculated (Asada et al.,1981; Ebert et al., 1984) relative effects of p- andd-electron contributions to the spin-lattice relaxationrate in this metal are smaller than in copper by a factorof 3. A similar reduction can be expected to hold forforces not mediated by s electrons.

3. Other interactions

In an interesting paper, Siemensmeyer and Steiner(1992) investigate the role of magnetoelastic coupling innuclear magnets, especially in copper. Such interactionsare important in electronic systems (Kotzler, 1984) butare usually assumed vanishingly small for nuclear spins.The authors find that, when spin-spin interactions areneglected, the quadrupolar energy transforms the fcccrystal of copper to a lattice with noncubic symmetry atTc'0.06 nK. The transition is similar to the Jahn-Tellereffect (Gehring and Gehring, 1975). Therefore, magne-tostrictive energy of copper is 2–3 orders of magnitudesmaller than spin-spin interactions. Siemensmeyer andSteiner emphasize, however, that their estimate is basedon conservative values of the shielding and antishieldingfactors. As a result, magnetostrictive energy could beeven 2 orders of magnitude higher, thus becoming sig-nificant in comparison with spin-spin interactions. Theauthors also discuss effects due to the external magneticfield and uniaxial stress caused by the weight of thesample itself. Effects due to distance dependence ofspin-spin interactions, i.e., exchange striction, are foundnegligible in comparison with magnetoelastic couplingwith quadrupolar forces.

Page 95: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

95A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Lindgard (1992) has studied the role of lattice dynam-ics in the calculation of exchange interactions. He findsthat zero-point lattice vibrations give rise to substantialcorrections in some of the exchange parameters, in com-parison with values calculated assuming a rigid lattice.The argument is based on the large separation of energyscales for the electronic, lattice, and nuclear systems.The nuclei see the average of phonon fluctuations,whereas the electronic system mediating the exchangeinteraction can adjust to them. Using the free-electronrange function for the RK interaction [see Eq. (64)],Lindgard estimates that the nearest-neighbor J1 is notmodified, whereas J2 and J3 are reduced by 20% and50%, respectively.

B. Mean-field theory of magnetic ordering

1. Basic equations

We rewrite the Hamiltonian of the spin system [seeEq. (1)] as

H52 12 (

i ,j

8IiAijIj2\gB•(

iIi , (73)

where the 333 matrix Aij consists of the dipolar andexchange interactions. The prime on the summation signindicates that the term i5j must be omitted. In themean-field theory, H is approximated by

HMF52\g(i

Bi•Ii112 (

i ,j

8^Ii&Aij^Ij&, (74)

where the local field Bi acting on spin i consists of theexternal field B and the field due to interactions withother spins, viz.,

Bi5B1(j

Aij^Ij&/~\g!. (75)

The field Bi is stationary, unlike the randomly fluctuat-ing local field B loc discussed in Sec. IV.C. The thermalaverage of a spin operator ^Ii& is given by

^Ii&5IBi

uBiuBIS \guBiu

kBT D , (76)

where BI(x) is the Brillouin function for spin I ,

BI~x !5@~I1 1

2 !/I#coth@~I1 12 !x#2~1/2I !coth~x/2!,

(77)

as in Eqs. (25a) and (25b).If the local fields are known, the energy E5^HMF& is

obtained from

E52 12 \gB•(

i^Ii&2 1

2 \g(i

Bi•^Ii&. (78)

2. Eigenvalue and other k-space equations

We first introduce the Fourier transforms

^I~k!&51N(

i^Ii&exp~2ik•ri!, (79)

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

A~k!5( 8j

Aijexp@2ik•~ri2rj!# . (80)

The energy per spin and the local field can then bewritten as

E/N52\gB•^I~k50 !&2 12 (

k^I~2k!&A~k!^I~k!&,

(81)

Bi5B1(k

A~k!^I~k!&cos~k•ri!/~\g!. (82)

Here we have assumed that the lattice has inversionsymmetry.

Further progress can be made by introducing the ei-genvalues ln(k) and eigenvectors en(k) of the Fourier-transformed interaction matrix. These are defined by

A~k!en~k!5ln~k!en~k!, (83)

where n51,2,3. Taking the en(k)’s as the basis vectors,we write

E/N52\gB•^I~k50 !&2 12 (

k,nln~k!an~k!2, (84)

an~k!5^I~k!&•en~k!. (85)

The local fields can be expressed in a similar way,

Bi5B1(k,n

ln~k!an~k!en~k!cos~k•ri!/~\g!. (86)

These equations provide a general and effective frame-work for carrying out the mean-field analysis.

3. Ordering vector and the ordering temperature

To obtain the critical temperature in zero field weseek solutions for small ^Ii& by linearizing Eq. (76). Thisis permissible for a continuous transition. One is led tothe equation (Kjaldman and Kurkijarvi, 1979)

l^I~k!&5A~k!^I~k!& (87)

with l53kBTc /I(I11). The problem thus reduces tosolving the eigenvalue equation (83). The physical solu-tion corresponds to the largest critical temperature Tc ,given by

kBTcMF5 1

3 I~I11 !lmax . (88)

Here lmax5maxk,n$ln(k)%, n takes values n51,2,3, andk runs over vectors in the first Brillouin zone. Theordering vector is the wave vector k for whichln(k)5lmax . The direction of the spins below T5Tc isdetermined by the eigenvector(s) corresponding tolmax .

At negative temperatures the situation is reversed.When T approaches zero from the negative side, thefirst solution to the linearized mean-field equation (87)is found at a wave vector that corresponds to the mini-mum of ln(k) (Abragam and Goldman, 1982).

The total amplitude of antiferromagnetic order at awave vector k can be obtained from a measurement of

Page 96: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

96 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

the antiferromagnetic Bragg reflection at this k. Apartfrom a constant coefficient, the structure factoruFAF(k)u25u^I(k)&u2 [see Eq. (51)].

4. Equal-moment and permanent spin structures

In equal-moment spin structures u^Ii&u is the same forall spins i . This is always the case at T50 as the spinsare saturated, provided that there is only a single spinspecies in the system. At T5Tc , however, Eqs. (87) and(88) determine the ordering vector irrespective ofwhether the moments u^Ii&u are equal or not. At in-between temperatures, the concept of permanent spinstructures (Villain, 1959; Abragam and Goldman, 1982)is useful. By definition, a spin structure is permanent ifthe local fields Bi satisfy

Bi5l^Ii&/\g , (89)

where l is independent of the site. If a spin structurecorresponding to the maximum of ln(k) can be chosenpermanent, l5lmax and the structure is stable withinthe mean-field theory, both at T50 and immediatelybelow T5Tc (Luttinger and Tisza, 1946; Villain, 1959;Abragam and Goldman, 1982; Kumar et al., 1986). Inthe intermediate-temperature region 0,T,Tc , thestructure is at least metastable, and its Gibbs free energyis lower than that of any other permanent configuration.

A permanent structure is always an equal-momentconfiguration. An equal-moment structure is permanentif and only if (i) the eigenvalues ln(k) for all nonzeroantiferromagnetic components ^In(k)&, kÞ0, are equaland (ii) the ferromagnetic component ^I(0)& dependson the external magnetic field B via ^I(0)&5\gB/(l2l(0)) (Viertio and Oja, 1993). From this itfollows that the critical field for a permanent spin struc-ture is

Bc5I@l2l~0 !#/\g . (90)

5. Thermodynamics

Thermodynamics is particularly simple for permanentspin structures. The partition function is given by

ZMF5expF2bN

2 S lI2p22\2g2B2

l2l~0 ! D GzIN~blpI !,

(91a)

where b5(kBT)21 and

zI~x !5sinh@~I1 1

2 !x#/sinh~ 12 x !. (91b)

The Gibbs free energy G52kBT lnZ. From theseequations one can derive all thermodynamic functions ofthe system.

The self-consistent equation for polarization is

p~T !5p~T ,B !5BI~blpI !, (92)

where the Brillouin function BI

was given by Eq. (77).Note that p does not depend on B anywhere in the an-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

tiferromagnetically ordered region. The temperature de-pendence of the critical field is

Bc~T !5p~T !Bc~T50 !, (93)

where Bc(T50) is given by Eq. (90).The expression for entropy per spin is

S52blI2p21lnzI~blpI !. (94)

Like p , S does not depend on B in the antiferromagneti-cally ordered (T ,B) region (Oja, 1984; Lindgard,1988a). This means, in particular, that the isentropes arevertical as shown in Fig. 7. Therefore, when the nuclear-spin system is demagnetized into the ordered state, thereis no further decrease in the spin temperature.13

When fluctuations, which are neglected in the mean-field theory, are taken into account, the above picturechanges slightly. It has been predicted that isentropesweakly bend towards higher temperatures when B→0 inthe ordered region (Lindgard, 1988a). Similar behaviorhas been observed in electronic magnets (Garrett, 1951;de Klerk, 1956). The absolute value of the ordered-phase entropy as predicted by mean-field theory is un-reliable. The critical entropy at B50 is equal to the en-tropy at T5` owing to neglect of short-range order. Atlow temperatures, on the other hand, the mean-fieldtheory neglects the entropy carried by spin-wave excita-tions.

The equation for longitudinal magnetic susceptibility(in SI units) is

x~T ,B !5m0r\2g2/@l2l~0 !# , (95)

which is constant in the ordered region, resulting in atemperature-independent magnetization M5xB/m0.

The longitudinal susceptibility in the ordered state,Eq. (95), can also be written as

x~T ,B !5C/@TcMF2uW# , (96)

where C is the Curie constant (see Table II), TcMF is the

mean-field ordering temperature [Eq. (88)], and theWeiss temperature uW5(R1L2D)C . Here R is thestrength of the exchange interaction [see Eq. (66)], L513 is the Lorentz constant, and D is the demagnetizationfactor along the direction of the external field (see Sec.IV.D). uW is related to the eigenvalue l(0), appearingin Eq. (95), by 1

3I(I11)l(0)5uWkB . Note that thehigh-T behavior of the susceptibility is given byx5C/(T2uW).

6. Transition from the polarized paramagnetic stateto the antiferromagnetic phase

Magnetic ordering was analyzed in Sec. XV.B.3 bylinearizing the mean-field equations at the ordering tem-

13The fact that the isentropes are vertical in the antiferromag-netically ordered state can also be seen from the thermody-namic relation (dB/dT)S5(]S/]T)B/(]M/]T)B by notingthat (]S/]T)B.0 always and that in the present case(]M/]T)B50 according to the mean-field theory.

Page 97: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

97A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

perature. The treatment was limited, however, to zeroexternal field. Although a similar calculation is possiblewhen BÞ0, we follow here another procedure.14 We de-scribe the transition from the polarized paramagneticstate to the antiferromagnetic phase using the soft-modetheory, which was first employed in this context by Lind-gard (1992). His approach was later elaborated by Ojaand Viertio (1993), who analyzed the general propertiesof a soft-mode transition in a system with anisotropicspin-spin interactions.

The idea of the soft-mode description of a phase tran-sition is the following: The energy needed to excite aspin wave at a wave vector k in the paramagnetic state isfirst calculated. At a field in which the excitation energyvanishes, i.e., when the spin wave becomes soft, theparamagnetic state is unstable with respect to antiferro-magnetic order characterized by this k vector. The or-dering vector Q=k is the one that becomes soft in thehighest field.

The starting point for the analysis is obtained from theearly work of Holstein and Primakoff (1940). In the po-larized state at T50, the excitation energy for a spinwave is

«k5ACk224uDku2, (97)

where

Ck52 12 I@Axx~k!1Ayy~k!#1\gB1IAzz~0 !,

Dk5 14 I@Axx~k!2Ayy~k!22iAxy~k!# . (98)

Here z is the direction of the external magnetic field andA(k) is the Fourier transform of the interaction matrix;see Eq. (80). The scheme for deriving this result is thefollowing (see, for example, p. 43 in Keffer, 1966): TheHamiltonian is rewritten using the spin-deviation opera-tors, higher-order terms than those bilinear in these op-erators are neglected, and the resulting Hamiltonian isFourier transformed and finally diagonalized. This pro-cedure yields the normal modes, i.e., spin waves and thecorresponding energies.

It is clear from Eq. (97) that in a high enough field allexcitation energies «k are positive. For antiferromag-netic interactions, one of the spin-wave energies eventu-ally becomes negative with decreasing B , indicating soft-ening of this particular excitation and instability of theparamagnetic phase.

To analyze the instability, it is convenient to write «kin the form (Oja and Viertio, 1993)

«k2/I25detS Axx~k!2\gB/I2Azz~0 ! Axy~k!

Axy~k! Ayy~k!2\gB/I2Azz~0 !D . (99)

The determinant vanishes when \gB/I1Azz(0) coin-cides with an eigenvalue l (k) of

14The soft-mode and mean-field theories lead to the sameprediction for an antiferromagnetic transition at T50 (Ojaand Viertio, 1993).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

A ~k!5S Axx~k! Axy~k!

Axy~k! Ayy~k!D . (100)

This 232 matrix is the xy block of the 333 matrixA(k) and corresponds to the plane perpendicular to thefield direction z. Therefore the soft-mode transitiontakes place at

Bc5I@ l max2l~0 !#/\g , (101)

where l(0)5Azz(0) and

l max5Maxk,n$l n~k!%. (102)

The wave vector k that yields l max is defined as Q . Notethe similarity between Eq. (101) and the relation validfor permanent spin structures, Bc5I@lmax2l(0)#/\g[see Eq. (90)].

The matrix A (k) clearly depends on the direction ofthe external field. Therefore the eigenvalues l n(k) canalso depend on the direction of B, and so can the order-ing vector Q and the critical field Bc .

A question of particular interest is whether the high-field ordering vector Q is the same as the zero-field or-dering vector Q which corresponds to the maximum ei-genvalue lmax of the 333 matrix A(k). The high-fieldsoft-mode transition has the following properties (Ojaand Viertio, 1993):

(i) If the largest eigenvalue lmax of A(k) is degener-ate, the soft-mode transition always takes place for theB50 vector Q. The critical field B5Bc [see Eq. (90)] isindependent of the field direction.

(ii) If the largest eigenvalue lmax of A(k) is nonde-generate, the soft-mode transition can take place for an-other ordering vector in addition to the one stable atB50, provided that the anisotropy is strong enough andthat its easy axis is not perpendicular to the field. By theeasy axis we mean the direction of the eigenvector cor-responding to lmax . The value of Bc will then be lowerthan that given by Eq. (90), since l max,lmax . However,if B is perpendicular to the easy axis, the high-field or-dering vector is the same as the one at B50, and Bc isnot lowered.

It should be noted, however, that if Bc is strongly sup-pressed by the anisotropy of the interactions, the transi-tion to the antiferromagnetic state may be of first order.Then the present approach, which is based on the as-sumption of a continuous transition, does not apply.

7. Ordering in the fcc lattice

Most features important for nuclear ordering in Cu,Ag, and Rh can be described in terms of a model thatcontains isotropic exchange interactions between near-est (J1) and next-nearest neighbors (J2) and dipolarinteractions between nearest neighbors @D15(m0/4p)\2g2r23]. Performing the lattice sums, we find

Page 98: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

98 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

A~k!5S J~k!24D1@2 12 cx~cy1cz!1cycz# 26D1sxsy 26D1sxsz

26D1sxsy J~k!24D1@2 12 cy~cx1cz!1cxcz# 26D1sysz

26D1sxsz 26D1sysz J~k!24D1@2 12 cz~cx1cy!1cxcy#

D ,

(103)

where

J~k!54~J122J2!~cxcy1cycz1czcx!

14J2~cx1cy1cz!226J2 . (104)

Here ca5cos(kaa), sa5sin(kaa), a5x ,y ,z , and the lat-tice constant is 2a . The eigenvalue equation can besolved analytically if the k vector lies in a symmetry di-rection of the reciprocal lattice (Oja and Viertio, 1993).

Let us first consider the case of only isotropic inter-actions, i.e., D150. The eigenvalue is then simplyl(k)5J(k) and the stable structure is the one that maxi-mizes J(k). Four different kinds of configurations areobtained; these are the principal ordering modes of thefcc system. One of them is the ferromagnetic structure;the three others are antiferromagnetic configurationswhose magnetic unit cells are shown in Fig. 110, illus-trated for the case when there is only one ordering vec-tor (single-k structure). It is also possible that the spinconfiguration is modulated by a superposition of cubic-symmetry-related vectors (multiple-k structure). Figure110 shows the unit cell for type-IV ordering as well. Itcannot be obtained from the J1,J2 model but becomesstable if the nearest-neighbor interaction is sufficientlyanisotropic.

The stability regions of the four principal modes ofordering in the J1J2 plane are shown in Fig. 87 (Smart,1966). The J1 /J2 ratios appropriate for Cu, Ag, and Rhare also indicated. The theoretically calculated values ofJ1 and J2 (Lindgard et al., 1986; Harmon et al., 1992)

FIG. 110. Magnetic unit cells of the four principal antiferro-magnetic configurations in fcc lattices. The respective orderingvectors are (I) k5(p/a)(1,0,0), (II) k5(p/a)( 1

2, 12, 1

2), (III)k5(p/a)(1, 1

2,0), and (IV) k5(p/a)( 12, 1

2,0). The nuclear-spindirections are opposite in sites marked by s and d , respec-tively.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

have been used for Cu and Ag, with all other interac-tions, including the dipolar force, being neglected. ForRh the experimental J1 and J2 were adopted (Hakonen,Vuorinen, and Martikainen, 1993). Type-I ordering ispredicted for the three metals at positive temperatures,as illustrated by the open symbols. At negative tempera-tures, shown by solid symbols, ferromagnetic order wasobtained.15 However, one should note that, at T,0, Rhis relatively close to the boundary between ferromag-netic ordering and antiferromagnetism of the secondkind. Experiments have shown, in fact, a tendency to-wards antiferromagnetism in rhodium at T,0 (see Sec.X.B).

The eigenvector equation (87) is trivial for the isotro-pic model but important when there is some anisotropy.In the case of type-I order, the dipolar energy forces^I(k)&, where k is a type-I ordering vector, to the planeperpendicular to k. Thus type-I systems have easy-planeanisotropy. In contrast, the dipolar energy chooses easy-axis anisotropy for the (0 2

323) order. Apart from deter-

mining the direction of ^I(k)& with respect to k, aniso-tropic interactions can also be decisive in the selection ofthe ordering vectors.

The principal kinds of antiferromagnetic structures offcc systems (see Fig. 110) can easily be chosen perma-nent when B50. This is because the phase factorcos(k•ri), where k is the corresponding ordering vector,takes only values 61 for the four types of orderings.[For types II, III, and IV the approriate phase factor isA2cos(k•ri1p/4)]. The situation is simple, however,only for the single-k structures depicted in Fig. 110 atB50. In an external field, anisotropic spin-spin interac-tions in some cases make it impossible to form perma-nent spin structures (Oja and Viertio, 1992). Therequirement of permanency also imposes severe restric-tions on multiple-k spin configurations.

C. Comparison of measured and calculated magneticproperties of Cu, Ag, and Rh

1. Mean-field Tc and the ordering vector

Kjaldman and Kurkijarvi (1979) were the first to in-vestigate nuclear ordering in a system with dipolar andRK interactions. They assumed the free-electron formfor the RK term, Eq. (64), and made their calculationfor different strengths of the exchange force by multiply-

15Of course, at T,0 the exchange constants are the same asat T.0. The signs of J1 and J2 have been reversed in Fig. 87 atT,0 because max$J(k)%5min$2J(k)%.

Page 99: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

99A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

ing the free-electron Jij’s with an overall scaling factor.16

Figure 111 reproduces their results. For the experimen-tal R520.42, as determined from the NMR measure-ments by Ekstrom et al. (1979), Tc

MF5230 nK. ForR520.37, which would reproduce the magic-angle spin-ning NMR measurement of ( jJ ij

2 by Andrew et al.(1971), Tc

MF would be 210 nK. Both mean-field pre-dictions are thus clearly higher than the experimentalvalue TN558 nK. The predicted ordering vector Q5(p/a)(1,0,0) as observed in both the low- and high-field regions (see Fig. 3).

If data from the band-structure calculations by Lind-gard, Wang, and Harmon (1986) are used for the RKinteraction, type-I order is again found with Tc

MF5181nK. These authors emphasized, however, that copper

16The coefficient h of Kjaldman and Kurkijarvi (1979) is re-lated to the R parameter through R520.2630.587h [Eqs.(64) and (66)].

FIG. 111. Magnetic ordering of Cu nuclei as a function of thestrength of the RK interaction, which is assumed to havethe free-electron form of Eq. (5). Upper frame: Transitiontemperature. Lower frame: Ordering vector Q. For uRu.0.15the spin order has the form Q=(p/a)(q ,0,0), whereas Q5(p/a)(q ,q ,0) for uRu,0.15. However, at R'20.11, Q5(p/a)(0.5,0.4,0) is found. Modified from Kjaldman andKurkijarvi (1979) and Huiku et al. (1986).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

falls in the region where the RK and dipolar interactionsstrongly compete. That is, if their theoretical Jij’s werescaled by 20%, from R520.34 to R520.27, the order-ing vector would move from Q5(p/a)(1,0,0) to an in-commensurate Q5(p/a)(h ,h ,0) with h'0.6, close tothe observed h5 2

3 in low and intermediate fields. Thisbehavior is illustrated in Fig. 112, which shows, in theform of a contour map, ln(k) on two high-symmetryplanes in the reciprocal space. There can be interestingfluctuation effects in a situation when there is, in addi-tion to the global maximum lmax , a nearly degeneratelocal maximum at another k vector, as in Fig. 112(a).

When anisotropic exchange forces and relativisticcorrections to the RK interaction are included, ln(k)becomes almost flat along the [100] directions (Oja,Wang, and Harmon, 1989). lmax is found at Q5(p/a)(0.87,0,0), but ln(k) is only 0.6% lower atQ5(p/a)(1,0,0). Anisotropic exchange interactionsthus have a tendency to destabilize type-I ordering. Thedifference for Q5(p/a)(h ,h ,0) is rather small as well,on the order of 10%. Anisotropic interactions alsoslightly suppress the ordering temperature becauseTc

MF is decreased to 170 nK, although the absolute valueof the exchange constant is increased to R520.37. Afurther decrease in Tc should result from increased spinfluctuations along the @j00# direction in k space.

Experiments (Poitrenaud and Winter, 1964; Oja, An-nila, and Takano, 1990; Hakonen and Yin, 1991) andtheoretical calculations (Harmon et al., 1992) all show astrong dominance by exchange forces in silver, withR'22.5. According to the free-electron picture [seeEq. (5)], the exchange constants Jij of Cu and Ag shouldbe equal, except for a constant coefficient. As a result,type-I ordering was expected (Oja and Kumar, 1987) forsilver on the basis of the mean-field calculations byKjaldman and Kurkijarvi (1979) on copper. Harmonet al., (1992) arrived at the same conclusion using their

FIG. 112. Contour plots for the eigenvalue maxn$ln(k)% [seeEq. (83)] of spin-spin interactions in two high-symmetry planesin the reciprocal space of the fcc lattice in copper. The RK-interaction strength parameter R520.26 and 20.42 in figures(a) and (b), respectively. Maximum eigenvalues were foundalong the GK direction at k'(p/a)(0.6,0.6,0) in (a) and at theX point k5(p/a)(1,0,0) in (b). From Lindgard et al. (1986).

Page 100: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

100 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Rev. Mod. Phys

TABLE VI. Summary of mean-field predictions for type-I ordering in Cu, Ag, and Rh at positivespin temperatures, and the corresponding experimental values. The quantities listed are the Neeltemperature TN [Eq. (88)], the critical field Bc(T50) at zero temperature [Eq. (90)], and the (lon-gitudinal) susceptibility x of the antiferromagnetic state [Eq. (95)]; x is independent of T and Bwithin the mean-field theory. The experimental x refers to its value at T5TN and B50 for Cu andAg, while for Rh we have listed the value at the lowest temperature T5280 pK reached so far. x andBc are given for zero demagnetization factor.

Cu Ag Rhtheorya exp.b theoryc exp.d theorye exp.f

TN (nK) 170 58610 2.34 0.5660.06 1.5 <0.2860.06Bc(T50) (mT) 417 270610 140 100610 101x (SI units) 2.94 3.0 0.321 0.3360.06 0.404 >0.55

aCalculated for exchange constants of Oja, Wang, and Harmon (1989).bFrom Huiku et al. (1986).cCalculated for exchange constants of Harmon, Wang, and Lindgard (1992).dFrom Hakonen and Yin (1991).eCalculated for exchange constants of Hakonen, Vuorinen, and Martikainen (1993).fFrom Hakonen, Vuorinen, and Martikainen (1993).

band-structure calculation of Jij . These predictions wereconfirmed by the neutron-diffraction measurements ofTuoriniemi, Nummila, et al. (1995).

At negative temperatures, ordering should proceedinto a ferromagnetic state that corresponds to lmin(Abragam and Goldman, 1982; Viertio and Oja, 1992).This state will be discussed in Sec. XV.H.

2. Bc(T50) and x of the ordered state

Table VI summarizes the mean-field predictions andexperimental values of TN , Bc(T50), and x for Cu, Ag,and Rh. The calculations were made for type-I orderingat T.0. The exchange parameters used for Cu and Agwere taken from first-principles electronic band-structure calculations (Oja et al., 1989; Harmon et al.,1992). For Rh the experimental J1 and J2 values wereused (see Fig. 87) (Hakonen, Vuorinen, and Marti-kainen, 1993).

The mean-field prediction for TN is in clear disagree-ment with the experiments. This is largely explained byinadequacies of the mean-field theory and will be dis-cussed in Sec. XV.C.3. For predicting magnetic proper-ties such as Bc or x , the mean-field theory is expected towork much better. The longitudinal susceptibilityxL5m0(dM/dB) has been measured in the orderedstate only in the case of copper (see Fig. 52), which al-lows a direct comparison with xMF. The observed xL isnot strictly constant as the mean-field prediction re-quires, but neither is the ordering only of type I. Varia-tions about the mean result are 614%. The value listedin Table VI is the zero-field susceptibility. In view ofthese ambiguities the excellent agreement betweentheory and experiment is partly fortuitous.

Also in the case of Ag, xMF agrees well with xexp atB50. Although xL has not been measured as a functionof the magnetic field, the nearly constant value of thetransverse susceptibility at TN (see Fig. 13 in Hakonenand Yin, 1991) suggests that xL behaves similarly aswell. That is, according to the mean-field theory as ap-

., Vol. 69, No. 1, January 1997

plied to a single-k type-I structure, the experimental sus-ceptibility along the external field is equal to x mea-sured perpendicular to the antiferromagnetic amplituded. Although x along d should be smaller than x perpen-dicular to d, the difference appears only with decreasingtemperature and vanishes at T5TN, so that the suscep-tibility is isotropic. For the same reason multiple-k statesat T5TN should also display isotropic susceptibility ofthe same amplitude.

For both Cu and Ag, the theoretical Bc is larger thanthe value extrapolated from measurements at a finiteT . The discrepancy is at least partly caused by quantumfluctuations. Although the spin is larger in Cu, quantumeffects may reduce Bc(T50) as much in Cu as in Agowing to the larger dipolar anisotropy in copper.

Since the ordered state has not been reached inrhodium, we have included in Table VI the susceptibilitymeasured at the lowest temperature produced in the dis-ordered state (Hakonen, Vuorinen, and Martikainen,1993); x(T5TN) should be larger. While the agreementbetween experimental and theoretical values of x isgood for Cu and Ag, there is a clear discrepancy in Rh:The measured x is about 40% larger than the mean-fieldprediction for the ordered state.

One important feature of the data shown in Table VIis that the magnetic susceptibility for Cu is higher thanthat for Ag and Rh by almost an order of magnitude.This has nothing to do with the larger magnetic momentof Cu nuclei, as one might at first think. To understandthe behavior it is useful to write Eq. (95) in the form

xMF51/~ l 2R2L1Dz!, (105)

where the magnitudes of the dimensionless R andl 5l/(m0r\2g2) parameters are set by the dipolar en-ergy. If exchange interactions were negligible, xMF

would have a universal value dependent on the crystalstructure only but not on the magnitude of the momentor of the spin. If the exchange is large, xMF is reduced.

Page 101: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

101A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Therefore the reason for the relatively small x in Ag andRh is that uRu is large in these metals.

Quantum-mechanical corrections neglected in themean-field theory have the effect, at least when T50, oflowering the susceptibility to less than the value pre-dicted by Eq. (105) (Oguchi, 1960; White et al., 1993).The effect is particularly large for I5 1

2. For an antifer-romagnet with the sodium chloride structure, for ex-ample, the correction would be on the order of 20%.Therefore, in Ag as well, the good agreement betweentheory and experiment for x may be partly fortuitous.

3. Beyond the mean-field theory

In the first experiments on copper (Ehnholm, Ek-strom, Jacquinot et al., 1979), no unambiguous signs ofordering were observed down to 50 nK, although themean-field prediction for TN was 230 nK (Kjaldman andKurkijarvi, 1979). This discrepancy inspired a largeamount of theoretical work aimed at improving esti-mates of the ordering temperature as well as predictionsfor the ordered structure. In addition, detailed compari-sons were made with the measured S vs T and x vs Tcurves.

Meaningful calculations for the paramagnetic state inzero field are more difficult than those for the orderedstate in the sense that the mean-field theory completelyfails, for example, in predicting the entropy, becauseshort-range order is neglected. Any realistic calculationof thermodynamic properties in the paramagnetic stateat B50 requires, therefore, a fairly sophisticated ap-proach, especially in the interesting range near TN .Apart from mean-field calculations (Kjaldman andKurkijarvi, 1979), the various theoretical techniques em-ployed have included spherical model (SM; Kumar et al.,1980; Kjaldman et al., 1981; Lindgard et al., 1986; Har-mon et al., 1992), high-temperature expansion (HTE;Niskanen and Kurkijarvi, 1981), linked-cluster expan-sion (LCE; Niskanen and Kurkijarvi, 1983), and ananalysis of the eigenvalue spectrum @ln(k)] of the inter-action matrix (Niskanen et al., 1982).

a. Spherical model

It is useful to review some of the results of thespherical-model calculations (Berlin and Kac, 1952;Mattis, 1985) as applied to copper and silver (Kumaret al., 1980). These results provide insights into fluctua-tion effects that are ignored in the mean-field theory.The ordering temperature Tc in the spherical model isobtained from

I~I11 !5kBTc

1N(

k,n@lmax2ln~k!#21. (106)

As in the mean-field theory (see Sec. XV.B.3), orderingtakes place to a structure corresponding to lmax5maxk,n$ln(k)% through a second-order phase transi-tion. In the spherical model, Tc depends on the wholeln(k) spectrum, rather than on lmax only as in themean-field theory. Particularly important is the part ofthe eigenvalue spectrum near l5lmax . From Eq. (106)

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

one can see that if the maximum at l5lmax is flat the ksum becomes large and Tc decreases. In other words, ifthere are many low-energy excitations around theminimum-energy configuration, thermal fluctuations willsuppress the ordering temperature.

The susceptibility is given by (Kumar et al., 1980)

x5m0r\2g2/@l~T !2l~0 !# , (107)

where l(T) is obtained from

I~I11 !5kBT1N (

k,n@l~T !2ln~k!#21. (108)

In the high-temperature limit l(T)@lmax , and onefinds l(T)53kBT/I(I11), which yields the Curie be-havior. At T5Tc

SM , l(T)5lmax , and the sphericalmodel gives the mean-field susceptibility [see Eq. (95)]for the ordered state. The expression for the entropy is

S/NkB5ln~2I11 !

21

2N (k,n

lnI~I11 !@l~T !2ln~k!#

3kBT. (109)

Kumar and co-workers first employed the sphericalmodel to study the role of fluctuations in copper (Kumaret al., 1980; Kjaldman et al., 1981). It was found that thisconsiderably lowered TN from the mean-field estimate,even by more than 50%. In these calculations exchangeinteractions were described by means of the free-electron RK force.

Lindgard et al. (1986) have more recently madespherical-model calculations17 in which the RK interac-tion was taken from their theoretical electronic band-structure data (see Table IV). They also made compari-sons with experiments by multiplying the exchangeconstants by an overall scaling factor to study the depen-dence of the results on R , the strength of the RK inter-action. They found TN582 nK for R520.42. This isquite close to the observed TN=58610 nK. However, asthe authors point out, TN

SM is for a second-order transi-tion and should therefore be lower than the observedTN because fluctuations must induce a first-order phasechange above the predicted second-order transition toshow consistency with experiments.

Another point that makes the discrepancy betweenthe observed TN and TN

SM for R520.42 more serious isthat the spherical model has a tendency to overestimatefluctuation effects and therefore to underestimate theordering temperature. This feature is particularly pro-nounced for a nearest-neighbor Heisenberg antiferro-magnet in an fcc lattice (Heinila and Oja, 1993b). Thespherical model does not predict long-range order at allwhen T.0, although there is long-range type-I orderbelow TN50.45J1 /kB according to Monte Carlo simula-tions, which, in principle, are accurate (Minor and Gie-bultowicz, 1988; Diep and Kawamura, 1989; Heinila andOja, 1993b).

17The correlation theory used by these authors is identical tothe spherical model in the limit considered.

Page 102: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

102 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

The ordering temperature of Ag has been calculatedusing the spherical model (Harmon, Wang, and Lind-gard, 1992) by employing the Ruderman-Kittel ex-change constants from band-structure calculations. Theresult, TN

SM51.5 nK, with the transition to an antiferro-magnetic structure of type I, is significantly higher thanthe measured TN=560660 pK.

b. High-T expansions and other quantum-spin theories

High-temperature expansion (HTE) offers a system-atic way to improve the accuracy of predictions for thedisordered state. HTE can also account for the quantumnature of spins, which is described by the sphericalmodel in only a trivial way, similarly to the mean-fieldtheory. Pade approximations provide a way of continu-ing the HTE to lower temperatures and can be used toobtain a reliable estimate of the ordering temperature(Rushbrooke et al., 1974).

Pirnie et al. (1966) have reported an HTE analysis offerromagnetic and antiferromagnetic Heisenberg mod-els with nearest-neighbor and next-nearest-neighbor in-teractions. The seven leading terms in the high-T expan-sions for the uniform (k=0) as well as the appropriatestaggered (kÞ0)18 susceptibilities were calculated as afunction of I . Divergence of the staggered susceptibilityat T5TN is an indication of a second-order transition toan antiferromagnetic state modulated by the corre-sponding k vector. An important observation made byPirnie et al. was that TN(I)/TN

MF(I) is remarkably spinindependent.

For a system interacting through dipole-dipole andthe free-electron RK forces, Niskanen and Kurkijarvi(1981) have carried out the (staggered) susceptibility ex-pansion to the fifth term and the specific-heat expansionto the fourth term. Their results showed that the Neeltemperature of Cu, with R5 –0.42, is 180 nK, which isabout 20% lower than the mean-field estimate but stillclearly higher than the observed TN .

A compromise between the spherical model and HTEis the linked-cluster expansion (LCE) of Niskanen andKurkijarvi (1983). The technique can describe quantumspin, and it reduces to the spherical model in the high-density limit for classical spins. In comparison with HTEresults for classical and I5 1

2 Heisenberg models (Rush-brooke et al., 1974), the linked-cluster expansion pre-dicts a much lower ordering temperature. The advan-tage of LCE, in comparison with the exact HTE, is thatthe method is better adapted to systems with long-rangeinteractions. For Cu with R=−0.42, LCE predictsTN=120 nK (Huiku et al., 1986). Since all calculationssystematically show that TN decreases when uRu is re-duced, the main reason for the discrepancy between cal-culations and the observed TN=58610 nK seems to be

18Staggered susceptibility describes the response to a nonuni-form field B(r)5B exp(ik•r), whereas the ordinary suscepti-bility gives the response to a uniform, constant field.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

that R520.42 for Cu is too large in absolute magnitude,at least when the free-electron form of the RK interac-tion is assumed.

c. Monte Carlo simulations

The TN of Cu has also been obtained from MonteCarlo simulations of classical spins. Three different stud-ies have been reported. The major differences betweenthese computations are in the assumptions made aboutthe exchange interaction. Frisken and Miller (1988a) ob-tained TN=50 nK using their calculation for the RK in-teraction (Frisken and Miller, 1986) up to eighth nearestneighbors. Viertio and Oja (1989, 1990a) found TN=65nK for a two-nearest-neighbor model and TN=33 nK fora set of longer-range exchange constants, which wereslightly modified from those given by the band-structurecalculations of Oja, Wang, and Harmon (1989).

In the case of silver, Monte Carlo simulations by Vi-ertio (1990), which include both dipolar and exchangeinteractions up to eighth nearest neighbors, yield a tran-sition to the expected type-I structure at TN=500 pK(Oja and Kumar, 1987). This is in good agreement withthe measured TN=560660 pK, while the mean-field es-timate is considerably higher, TN

MF=2.3 nK.At negative spin temperatures, Monte Carlo simula-

tions (Viertio and Oja, 1992) again produce good agree-ment with experiment, TC

MC=21.7 nK while theobserved TC=−1.960.4 nK (Hakonen, Nummila, Vuo-rinen, and Lounasmaa, 1992). The RK interaction usedin these simulations on silver nuclei was obtained fromthe calculations of Harmon et al. (1992) and correspondsto R=22.3.

An important and difficult question is how the quan-tum nature of spins should be taken into account whenMonte Carlo results for classical spins are comparedwith experimental data. In the Monte Carlo resultsquoted above, no quantum corrections were made. Thesimulations were done for classical spins with magneticmoments m5g\I . An estimate of Tc for the correspond-ing quantum-spin system can be made using the resultobtained from the high-temperature expansions ofPirnie et al. (1966), according to which Tc(I)/Tc

MF(I) isnearly spin independent for Heisenberg systems in anfcc lattice. In other words, Tc scales approximately withI as predicted by the mean-field theory. Therefore theordering temperature Tc(I) for a system of quantumspins with moments g\I can be estimated from

Tc~I !5I11

ITc

MC , (110)

where TcMC is obtained from Monte Carlo simulations on

a system of classical spins with dipole momentsumu5g\I , using the same spin-spin interactions.

In the case of copper, I5 32, the Monte Carlo results

quoted above should thus be multiplied by 53, while for

silver the multiplier is 3. When this is done, there isalmost a factor-of-3 discrepancy between predicted andobserved ordering temperatures of silver, both at posi-tive and negative spin temperatures. At T.0, the

Page 103: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

103A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

(I11)/I-corrected TN then agrees with the spherical-model result of Harmon et al. (1992).

One should note, however, that the (I11)/I scaling,based on the work of Pirnie et al. (1966), is for a second-order phase change although the transition is of first or-der according to simulations (Viertio, 1992) and experi-ments (Hakonen and Yin, 1991). Since TN for a first-order transition is higher than for a correspondingsecond-order phase change, the remaining discrepancy isin fact less than a factor of 3. Nevertheless, it seems thatthere remains a clear disagreement between the theo-retical and measured ordering temperatures of silver.

When TN=33 nK from the Monte Carlo simulationson copper by Viertio and Oja (1990a) is scaled by(I11)/I , one obtains TN=55 nK, in excellent agreementwith the measured TN=58610 nK. Recent simulationsby Heinila and Oja (1995), which include couplings onlybetween three neighboring shells, yield TN=33 nK whenscaled by (I11)/I . The earlier simulations (Frisken andMiller, 1988a; Viertio and Oja, 1989) give a slightly toohigh TN when scaled by (I11)/I . Among these simula-tions, only the more recent calculations (Viertio andOja, 1990a; Heinila and Oja, 1995) are able to accountfor type-I as well as type-(0 2

323) spin configurations. It

seems, therefore, that the near degeneracy of the (1 0 0)and (0 2

323) orders enhances spin fluctuations and results

in a reduction of TN. This is consistent with spherical-model calculations by Lindgard et al. (1986) and is easyto understand: In the case of a near degeneracy, theeigenvalue ln(k) must be rather flat in the k space be-tween the k=(p/a)(1,0,0) and k=(p/a)(0, 2

3,23) positions,

which is illustrated in Fig. 112(a). As discussed in con-nection with Eq. (106), fluctuations of k vectors in thisregion of the reciprocal lattice will then lower TN .

Viertio and Oja (1992) have made Monte Carlo cal-culations to simulate the adiabatic demagnetization pro-cess of nuclear spins in silver using the method devel-oped by Merkulov et al. (1988).

4. S and x of the paramagnetic state

A comparison between the measured and calculatedsusceptibilities of copper was presented in Fig. 25. It isobserved that for R520.42, both the spherical modeland LCE predict a too-large inverse susceptibility. Thesituation is the opposite for the simple Curie-Weiss ap-proximation, which is included for comparison. Withinthe spherical model, the measured and calculated x vsT curves can be fitted well if uRu is changed toR520.28 (Lindgard et al., 1986). The RK interactionused in LCE was the free-electron approximation, whilethe band-structure values were employed in thespherical-model calculation. Unfortunately, however,comparisons between theory and experiment for the Svs T and x vs T curves are not yet available for the mostcomplete set of theoretical exchange interactions, whichalso include anisotropic coupling (Oja et al., 1989).

Figure 27 shows a comparison with the S vs T curve.Now the LCE prediction with R520.42 describes thedata well down to the ordering temperature. The

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

spherical-model estimate for S is too low. The agree-ment with the spherical model would improve if a lowerR were assumed, but agreement with the linked-clusterexpansion would then become worse.

Far less theoretical work has been done to computethe thermodynamic properties of silver above TN. Thesusceptibility of Ag, calculated using the sphericalmodel, agrees best with measurements if the strength ofthe RK interaction is decreased in absolute value fromR522.3 to R521.7 (Harmon et al., 1992). The S vsT curves of Ag and Rh have been investigated by Ha-konen (1994) using Monte Carlo simulations of classicalspins.

D. Fluctuation-stabilized type-I spin configurations

1. Continuous degeneracy of the mean-field solution

The most general spin configuration for the fcc type-Iorder can be written as a superposition of the threewave vectors k15(p/a)(1,0,0), k25(p/a)(0,1,0), andk35(p/a)(0,0,1),

^Ii&/I5m1 (j51,2,3

djcos~kj•ri!, (111)

where m is the magnetization, dj is the amplitude of theantiferromagnetic modulation, and kj is the correspond-ing wave vector. The structure can hence be a single-k,double-k, or a triple-k state, and the number of sublat-tices can vary from 2 to 4.

Dipolar anisotropy fixes the relative directions of djand kj by imposing the constraints (Luttinger and Tisza,1946)

dj•kj50, j51,2,3. (112)

The origin of this anisotropy can be seen from the eigen-value equation (87). For simplicity, let us consider theJ1J2D1 model of Sec. XV.B.7. From Eq. (103) one findsthat A(k1) is diagonal. The doubly degenerate eigen-value of the yz block, l5l25l3524J116J214D1, isthe relevant eigenvalue because l2l1512D1.0. Or-dering therefore takes place in the plane perpendicularto k1. The energy scale associated with the anisotropy islarge, on the order of the dipolar energy. Numerical cal-culations give l2,32l156.50(m0/4p)r\2g2 when the di-polar interactions are fully included (Cohen and Keffer,1955).

Requiring equal moments at the four sublattices, onefinds (Kumar et al., 1986)

umu21ud1u21ud2u21ud3u25p2, (113a)

m•d11d2•d350, m•d21d3•d150,

m•d31d1•d250. (113b)

Here p is the total sublattice polarization, u^Ii&u/I . Thestructure is permanent if

m~T ,B!5B/Bc~T50 !, (114)

where

Page 104: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

104 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Bc~T50 !5I@l2,32l~0 !#/\g . (115)

The antiferromagnetic order parameter can havealtogether six nonzero components (d1y ,d1z ,d2x ,d2z ,d3x ,d3y). The four equations, (113a) and (113b), donot fix the solution uniquely; there remains a continuoustwo-dimensional degeneracy. To single out the groundstate one has to consider fluctuation effects.

2. Static susceptibility matrices

Thermodynamics of type-I structures is given by thegeneral results for permanent spin structures, whichwere discussed in Sec. XV.B.5. In particular, the longi-tudinal susceptibility in the ordered state,

x0[m0r\2g2/@l2,32l~0 !# , (116)

is independent of T and B. The value of the full suscep-tibility matrix x(0) depends on the actual spin configu-ration. Calculation of x(0) is technically rather compli-cated. Special cases have been treated by Heinila andOja (1996).

We consider first a single-k structure in which themodulation vector k35(p/a)(0,0,1) is parallel to thefield aligned along a crystalline axis, say [001]. In thiscase

^Ii&/I5~0,0,m !1~0,d3,0!cos~k3•ri!, (117)

where m5B/Bc and m21d325p2. The static susceptibil-

ity is

x~0 !5S x0 0 0

0 xyy 0

0 0 x0

D , (118)

where

xyy /x05BDBc p2~12t!1B0

2lt/@l2,32l~0 !#

BDBc p2~12t!1~V' /g!2lt/@l2,32l~0 !#.

(119)

Here

BD5~I/g\!~l2,32l1! (120)

is the anisotropy field for type-I order, and t5t(p) is amonotonic function of polarization defined by the para-metric equation

t@p~x !#512x

p~x !

dp~x !

dx, p~x !5BI~x !. (121)

In particular, t assumes the values t(0)50 andt(1)51. At B50 we find the well-known behavior inwhich the susceptibility along the staggered magnetiza-tion (xyy) increases from zero at T50 to xyy5x0 atT5TN , whereas the susceptibility perpendicular to thespins is constant (Kittel, 1971).

The above case may be considered, however, as a spe-cial case. A more general single-k structure is obtainedin a finite field that is not parallel to the ordering vector.The spin configuration can be described by

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

^Ii&I

5B0

Bc1d

B03kuB03ku

cos~k•ri!, (122)

where m21d25p2. Here k is one of the three type-Iordering vectors, namely, (p/a)(1,0,0), (p/a)(0,1,0), or(p/a)(0,0,1). For this structure, Heinila and Oja (1996)have found a completely isotropic susceptibility. Moreprecisely, x(0)5x0I . The same x(0) has been found fora triple-k state, defined by Eq. (140). For more details,consult the original paper.

According to the above results for a single-k state,x(0) behaves in a discontinuous fashion when the struc-ture of Eq. (122) is demagnetized to zero field. It wassuggested that this is connected with a reorientationtransition in low fields, which is similar to a spin-floptransition.

3. Overview of fluctuation mechanisms

Several mechanisms can lift the continuous, two-dimensional degeneracy that the mean-field theory pre-dicts for the various type-I fcc antiferromagnets:

(i) Thermal fluctuations.(ii) Quantum fluctuations.(iii) ‘‘Quenched-in randomness’’ brought about by the

presence of different moments, in the case of atwo-isotope system.

(iv) Defects. They also act as quenched-in random-ness.

(v) Magnetoelastic forces.

Much theoretical work has been devoted to the firsttwo possibilities, the selection of the ground state byfluctuations. For type-I fcc antiferromagnets with isotro-pic interactions, the spin-wave calculation of Oguchiet al. (1985) first showed that quantum fluctuations favora single-k structure instead of, for example, a triple-kconfiguration. Effects due to dipolar anisotropy as wellas configurations caused by the external magnetic fieldwere first discussed by Viertio and Oja (1987), who alsoemployed spin-wave theory. The same problems werelater investigated by Lindgard (1988a, 1988b), using per-turbation theory, and more recently by Heinila and Oja(1993a). The effects of thermal fluctuations have beenstudied by employing Monte Carlo simulations (Friskenand Miller, 1988a; Frisken, 1989; Viertio and Oja, 1989;Viertio, 1990), spin-wave theory (Viertio and Oja, 1989),Ginzburg-Landau theory (Oja and Viertio, 1987), andthermodynamic-perturbation theory (Heinila and Oja,1994a).

The multitude of various studies reflects the delicatenature of the problem. In the case of copper, an addi-tional complication is caused by competition betweenthe dipolar and RK interactions. There were several dis-crepancies among the early theoretical investigations,but now it seems that recent studies by Heinila and Oja(1993a) have clarified the situation. For this reason weshall focus mainly on their work. We shall first discussthermal fluctuations, which are intuitively more obviousthan the corresponding quantum phenomena. It turns

Page 105: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

105A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

out that, at least to the leading order, thermal and quan-tum fluctuations stabilize the same spin structures.

Before we continue, however, we shall briefly com-ment on the last three possibilities. Following the discus-sion by Henley (1987), an ad hoc argument has beenproposed according to which the presence of two iso-topes in Cu and Ag is less important for the selection ofthe ground state than, for example, quantum fluctua-tions (Oja and Viertio, 1987). Systematic Monte Carlosimulations so far have not been performed to investi-gate this question.

In copper, with I5 32, the most important consequence

of various defects is probably the quadrupolar interac-tion that arises if there are distortions from cubic sym-metry. The strength of the quadrupolar force near sev-eral substitutional impurities has been investigated byNMR experiments (Rowland, 1960). In heat-capacitymeasurements of high-purity polycrystalline copper atmillikelvin temperatures, an extra contribution to theheat capacity of a pure system has been observed andattributed to quadrupolar interactions near impuritiesand lattice imperfections (Gloos et al., 1988). One canconclude, however, that at least the salient features ofnuclear ordering in copper are not related to defectsbecause the main characteristics of the data do not de-pend on a particular sample.

Siemensmeyer and Steiner (1992) have investigatedthe importance of magnetoelastic coupling betweennuclear spins and the lattice. They find that magneto-elastic energy due to quadrupolar forces can be signifi-cant in the selection of the ground state among degen-erate type-I structures.

4. Thermal fluctuations

Heinila and Oja (1994a) have recently investigatedthe selection of the ground state in type-I fcc antiferro-magnets by thermal fluctuations. They made use of theso-called thermodynamic perturbation theory (Landauand Lifshitz, 1980), which, in this context, is identical tothe spin-wave theory for classical spins. Although thisapproach is limited to low temperatures, it is valuablebecause it gives a simple expression for the leadingground-state selection term.

In order to describe fluctuations about the T50 con-figuration for a spin at site i , it is useful to introduce alocal coordinate frame (ex

i ,eyi ,ez

i ) where ezi is parallel to

the equilibrium direction. The spin vectors can be writ-ten as

Ii5I@ezi cosu i1sinu i~ex

i cosf i1eyi sinf i!#

5ezi ~I2ua iu2!1~I2ua iu2/2!1/2~a i* e1

i 1a ie2i !, (123)

where the complex quantities a i are defined in terms ofthe polar angles as a i5I1/2eif i(12cosui)

1/25ui1iv i ande6

i 5221/2(exi 6iey

i ). This is the classical analog of theHolstein-Primakoff transformation. The usefulness ofEq. (123) relies on the fact that sinuidfidui } duidvi . Re-taining only the parts bilinear in a i and a i* , we rewritethe Hamiltonian of Eq. (73) as H5H01H1 with

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

H05E01lI(i

a i* a i , (124)

H152 12 I(

i ,j

8$e1

i Aije1j a i* a j* 1e1

i Aije2j a i* a j1c.c.%,

(125)

where E052 12NlI2 and l is related to the site-

independent local field through B loc5lI/(\g) [see Eq.(89)]. The Gibbs free energy can be expanded with re-spect to H1 in the form

G'G02^H12&/~2kBT !, (126)

where G0 is the free energy associated with H0 and

^H12&}E )

i@duidv iexp@2lI~ui

21v i2!/kBT#H1

2 ,

(127)

up to a normalization factor. Extension of integrationover the whole complex plane also includes the unphysi-cal states ua iu2.2I of the conventional spin-wave theo-ries. Owing to this defect the analysis applies to the low-temperature region only. One obtains (Heinila and Oja,1994a)

G5E02TS02NkBT ln~T/T0!2kBT

2l2

3(i ,j

8$ue1

i Aije1j u21ue1

i Aije2j u2%, (128)

where S0 and T0 are constants. This result is an estimateof the lowest-order spin-wave free energy. When theclassical ground state at T50 has infinite degeneracy,Eq. (128) can be used to find the configuration favoredby thermal effects.

5. Quantum fluctuations

Quantum fluctuations can be treated in the same wayas thermal fluctuations. Now the variables a i* and a i ofthe previous section simply become the bosonic creationand annihilation operators for spin deviations, ai

† andai . In the second-order perturbation theory, the energyat T50 is E5E01DE , where (Heinila and Oja, 1993a)

DE52I

4l(i ,j

8ue1

i Aije1j u2. (129)

This expression is only slightly different from the cor-responding term causing ground-state selection by ther-mal fluctuations, Eq. (128). Heinila and Oja (1994a)have found that, apart from unimportant constants, theground-state selection is the same for thermal and quan-tum fluctuations if the interactions are isotropic. Thesame may also be true when the interactions are aniso-tropic, but a rigorous proof is lacking. At least in allspecial cases investigated by Heinila and Oja it wasfound that quantum and thermal fluctuations favor thesame spin structures.

Page 106: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

106 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Perturbation theory is less accurate than the linearspin-wave theory which was used previously to discussthe ground state of nuclear magnets (Viertio and Oja,1987). The transparency of perturbation theory, to-gether with the fact that nearest-neighbor correlationsare by far the most important for lifting the degeneracy,makes it a reasonable approach to the ground-stateproblem, as has been emphasized by Lindgard (1988a).

6. Isotropic spin-spin interactions

Before discussing the ground-state spin configurationsof anisotropic fcc antiferromagnets, it is useful to con-sider the isotropic Hamiltonian,

H52 12 (

i ,j

8JijIi•Ij2\gB•(

iIi. (130)

In this case the direction of the magnetic field with re-spect to the crystalline axes has no influence.

The free energy of Eq. (128), describing thermal fluc-tuations, becomes (Heinila and Oja, 1994a)

G5E02TS02NkBT ln~T/T0!2kBT

4l2

3(i ,j

8Jij

2 @11I24~Ii•Ij!2# , (131)

while the correction to the ground-state energy due toquantum fluctuations is (Long, 1989; Larson and Hen-ley, unpublished; Heinila and Oja, 1993a)

DE52NIJ1

2

8l F10216m21I24 (a.b

~Ia•Ib!2G , (132)

where a and b denote summation over the four sublat-tices. These two expressions show that both quantumand thermal fluctuations favor the spin configurationthat maximizes ( i ,j(Ii•Ij)

2. Thus fluctuations tend to sta-bilize collinear spin arrangements.

The ground-state selection term ( i ,j(Ii•Ij)2 was, in

fact, first proposed by Henley (1987, 1989), who used asomewhat different derivation. His interpretation of theresult is that fluctuations of a fixed-length spin j gener-ate random exchange fields dBi5JijdIj which, to leadingorder, are perpendicular to the equilibrium direction ofIj . If the equilibrium spin structure is collinear, dBi isperpendicular to ^Ii&. Spin i can then lower its energy byrelaxing towards the direction of the fluctuating field.

The evolution of the spin structure with the externallyapplied field has been investigated by Heinila and Oja(1993a). The results are illustrated in Fig. 113. In zerofield the stable structure is a single-k state with sublat-tice spin directions I15I252I352I4. At low B , spinscant towards the field to match m5B/Bc as required byEq. (114). There is a transition at m5B/Bc50.407 froma single-k configuration to a triple-k state. In the latterstructure at 0.407,m,0.5 [see Fig. 113(b)], spins in thethree sublattices are tilted towards the field so that twoof them have the same spin direction; the spins in thefourth sublattice are nearly opposite to the external

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

field. At B5B/Bc50.5, a collinear structure is againpossible, with I15I25I352I4 and parallel to B. In thetriple-k structure at m>1/2, depicted in Figs. 113(c) and113(d), there are three sublattices in which spins havethe same direction. In the triple-k states of Figs. 113(b)–113(d), the spins in the four sublattices are always copla-nar.

The evolution of the ground state with an externalmagnetic field has also been investigated by MonteCarlo simulations (Heinila and Oja, 1994b) which, inprinciple, yield the exact solution for classical spins.These simulations were made for the Hamiltonian ofEq. (130) assuming an antiferromagnetic nearest-neighbor exchange interaction, J1,0, and a weak ferro-magnetic next-nearest-neighbor interaction J2520.1J1.The spins were taken as unit vectors. The results werecharacterized by the parameters

p'2 5(

j51

3

$~dj!x21~dj!y

2%, (133a)

p i25(

j51

3

~dj!z2 , (133b)

where the z axis is along B. The behavior of p'2 and p i

2

at T50.6uJ1u/kB is illustrated in Fig. 114 for two simula-tions, one for an increasing and the other for a decreas-ing field. The simulated sample consisted of N54096spins, and periodic boundary conditions were imposed.

From the figure one can identify the low-fieldstructure of purely transversal modulation. At B↑'5uJ1u/(\g), a discontinuous transition takes place to acollinear configuration with purely parallel antiferro-magnetism; note the quite large hysteresis. AtB'6uJ1u/(\g), there is a continuous transition to astructure with mixed parallel and perpendicular modu-lations. Finally, the critical field to the paramagneticstate is Bc'8uJ1u/(\g) at TN50.6uJ1u/kB . The low-field

FIG. 113. Isotropic, type-I fcc antiferromagnet in an externalmagnetic field. Directions of the spins in the four sublatticesI1–I4 are indicated. d1–d3 are the corresponding amplitudes ofthe antiferromagnetic modulation, and m is the magnetization[see Eq. (111)]. The external field points upwards. (a)B50.40Bc ; (b) B50.41Bc ; (c) B50.5Bc ; (d) B50.9Bc .From Heinila and Oja (1993a).

Page 107: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

107A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

perpendicular phase can be identified as the structure inFig. 113(a) and the parallel configuration as the up-up-up-down state in Fig. 113(c). The behavior of paralleland perpendicular modulations in high fields are inagreement with Fig. 113(d).

These simulations were carried out at a temperatureT50.6uJ1u/kB , which is very close to the correspondingzero-field Neel temperature TN50.68uJ1u/kB . A low-Tground-state analysis, based on maximizing ( i ,j(Ii•Ii)

2,therefore works remarkably well in a region where onemight expect changes from the low-T behavior.

7. Anisotropic spin-spin interactions with an easy plane

Ground-state selection has been investigated quite ex-tensively for anisotropic models because of its relevanceto nuclear ordering, particularly in copper. The firsttheoretical studies were made by Viertio and Oja (1987),who used the linear spin-wave theory to calculate thequantum-mechanical correction to the ground-state en-ergy, viz.,

E5E01(k

FEk1(a

12 \vk,aG , (134)

where Ek is a structure-dependent term, \vk,a is thespin-wave energy, and a is an index of the mode. TheHamiltonian consisted of the complete, infinite-range di-polar interaction and the free-electron RK coupling.Magnetic phase diagrams were calculated as a functionof B , the direction of B with respect to the crystallineaxes, and R , the strength of the RK interaction. Owingto the extensive computing time requirements it was notpossible to perform a complete search of the groundstate in the whole degenerate, two-dimensional param-eter space. Instead, an ansatz for the ground state wasmade by choosing trial spin configurations of particu-

FIG. 114. Field evolution of an isotropic, type-I fcc antiferro-magnet, as given by Monte Carlo simulations. Total parallelp i

2 (s) and transversal p'2 (d) antiferromagnetic modulations

at T50.6uJ1u/kB are shown. The arrows indicate directions offield changes. From Heinila and Oja (1994b).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

larly high symmetry with respect to B and the crystallinedirections. The B50 spin structures of Viertio and Ojaare shown in Fig. 115. For strong exchange, the groundstate is a single-k configuration as was already found byOguchi et al. (1985). When the dipolar interaction is suf-ficiently strong, i.e., when uRu,0.55, the ground state is adouble-k structure.

The problem was further investigated by Lindgard(1988a, 1988b), who introduced the perturbation ap-proach for studying quantum fluctuations. The methodwas technically simpler than the spin-wave calculationand yet contained the essential physics. Another usefulsimplifying approximation of Lindgard was to focus theanalysis on the nearest-neighbor interaction, which isthe most important quantity in the ground-state selec-tion. At B50, his results agreed with spin-wave calcula-tions. Considerable differences between the two ap-proaches were found, however, in an external field. Thisobscured the theoretical situation although Lindgard’sresults provided a possible explanation for the threephases observed in the susceptibility measurements of acopper single crystal (Huiku et al., 1986).

Heinila and Oja (1993a) have recently repeated theperturbation calculation of the ground state using thesame approximation as Lindgard. The two sets of resultswere clearly different for Bi@110# and in the high-fieldregion for Bi@001# . Since a careful examination (Heinilaand Oja, 1993a) of the discrepancies clearly shows thatLindgard’s results are partly erroneous, we have chosento present here only the phase diagrams of Heinila andOja.

Monte Carlo simulations have also been performed toinvestigate the magnetic phase diagrams of copper andsilver (Frisken and Miller, 1988a; Frisken, 1989; Viertioand Oja, 1989; Viertio, 1990). These results are valuableas they provide, in principle, an accurate result for theground state of classical spins. The delicate nature of theproblem presents several numerical difficulties. The re-laxation times in the simulations can be very long, asthere are typically several local minima of free energy.The long range of the interactions also makes the calcu-lations demanding in terms of CPU time. These reasonshave strongly limited the range of fields and exchangeconstants that have been investigated numerically.

FIG. 115. Ground-state spin configurations in zero externalmagnetic field assuming dipolar and free-electron RK interac-tions: (a) uRu,0.55; (b) uRu.0.55. From Viertio and Oja(1987).

Page 108: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

108 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

a. Bi[001]

The phase diagram resulting from the quantum-mechanical perturbation calculation by Heinila and Oja(1993a) is shown in Fig. 116. The coordinates in the fig-ure are the magnetization m5B/Bc and the parameterd53D1 /(D12J1) expressing the strength of thenearest-neighbor dipolar interaction D15(m0/4p)\2g2r23 over the exchange force J1. All possibletype-I spin structures with easy-plane anisotropy [seeEqs. (112) and (113)] were included in the calculation.The diagram contains four different phases. They arelabeled according to the number of nonzero dj vectors.The single-k and double-k phases 1a, 1b, and 2 are de-scribed by spin projections on a plane perpendicular toB. The effect of the magnetic field is only to alter thelength of every spin projection in the same way. For atriple-k structure, however, the projection diagram candepend on B . The descriptions of the triple-k states 3 inFig. 116 are, therefore, only schematic.

The spin configurations were obtained from Eq. (111)where the vectors dj have the following expressions:

1a: d35pAF

A2~ex1ey!, (135a)

1b: d15pAFey . (135b)

2: H d15~pAF /A2 !ey ,

d25~pAF /A2 !ex .(135c)

3: H d15pAFsinu cosfey ,

d25pAFsinusin fez ,

d35pAFcosuey .

tanf52pAF

mcosu , (135d)

FIG. 116. Phase diagram for easy-plane type-I fcc antiferro-magnets in a magnetic field oriented along the [001] crystallinedirection. Horizontal axis is the magnetization m5B/Bc , andthe vertical axis is d53D1 /(D12J1), where D1 and J1 are thenearest-neighbor dipolar and exchange interactions, respec-tively. The various spin configurations have been illustrated byprojection diagrams with labels indicating the number ofk-vectors in the structure. From Heinila and Oja (1993a).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Here pAF5A12m2.Along the lines d50, d51, and m51/2 the triple-k

state 3 takes a particularly symmetric form,

H d15d35@m~12m !#1/2ey ,

d25~m21 !ez .(136)

The spins in the four sublattices are [see Eq. (111)]

5I1 /I52@m~12m !#1/2ey1~2m21 !ez ,

I2 /I5ez ,

I3 /I522@m~12m !#1/2ey1~2m21 !ez ,

I45I2 ,

(137)

where the subscripts 1–4 refer to Fig. 117. Spins in twoof the sublattices are parallel to the field, whereas spinsin the other two tilt towards the field when B is in-creased. This structure is the intermediate-field triple-kstate found earlier by Lindgard (1988a).19 One shouldnote, however, that far from the lines d50, d51, andm51/2 phase 3 differs considerably from this symmetricform. Moreover, a separate analysis must be performedin the limit d50, as will be discussed in the context ofFig. 119.

The results in Fig. 116 are in fairly good agreementwith the earlier spin-wave calculation of Viertio and Oja(1987), except for structure 3, which was not investi-gated in their work. In comparison with the earlier per-turbation calculation of Lindgard (1988a) there is goodoverall agreement for m,0.8 except for the fact that thetriple-k structure does not generally have the symmetricform of Eq. (137), as he found. In high fields, m.0.8,there are considerable differences. This field region isparticularly important because of the possibility of mak-ing comparisons with neutron-diffraction data on cop-per.

Frisken and Miller (1988a) have performed MonteCarlo simulations to investigate the magnetic phase dia-

19Lindgard used the symbol o↓

o↑

for the structure.

FIG. 117. Enumeration of the four sublattices in type-I fccantiferromagnets. Spins ^Ii& at sites i with the same label areequal. The spin vectors illustrate a single-k structure.

Page 109: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

109A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

gram of copper. A special effort was made to obtain asreliable results as possible. The long-range dipolar andexchange interactions were included up to eighth-nearest neighbors, thereby considering 140 spins. Theindirect exchange interaction was obtained from theirprevious electronic band-structure calculation (Friskenand Miller, 1986). These authors found the ordered spinconfigurations as a function of B , applied along the [001]crystalline axis. Their results are shown in Fig. 118.There are four separate spin configurations, denoted byAF1–AF4. The low- and high-field phases are two-sublattice structures, whereas AF2 and AF3 are four-sublattice states. The division of the total antiferromag-netic order into transverse and longitudinal componentsseemed compatible with spin arrangements deducedfrom static susceptibility measurements of a coppersingle crystal (Huiku et al., 1986). Later neutron-diffraction experiments showed, however, that in lowand intermediate fields copper nuclei display (0 2

323) or-

der as well.Nevertheless, it is interesting to compare the results of

Frisken and Miller (1988a) with the calculations by Hei-nila and Oja (1993a) for the type-I structure. The order-ing vector in the single-k configurations AF1 and AF4 ofFrisken and Miller is along the field, as it is for state 1aof Eq. (135a). AF2 falls in the general class of triple-kstates [see Eq. (135d)], while AF3 refers to the symmet-ric triple-k spin configuration of Eq. (136). Since thenearest-neighbor interaction assumed by Frisken and

FIG. 118. Predicted spin order for copper in different appliedmagnetic fields as given by the Monte Carlo simulations ofFrisken and Miller (1988a). In all cases the ordering is confinedto the yz plane.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Miller corresponds to d51.59, the perturbation calcula-tion would yield, according to Fig. 116, the evolutionscheme 1a→3→1b with increasing field. Thus the high-field structure and the detailed form of the triple-k stateare different.

Viertio and Oja (1989) have also made Monte Carlosimulations of nuclear ordering in copper. Dipolar inter-actions with nearest and next-nearest neighbors wereincluded in addition to exchange interactions withJ1 /kB=212.5 nK and J2 /kB=5.6 nK. Three structureswere obtained with increasing field, applied along the[001] crystalline axis, namely, 1b, 2, and 1b. Since thenearest-neighbor interaction used in these simulationscorresponds to d52.01, there is again some but not com-plete agreement with the perturbation calculation of Fig.116. As the d parameter of copper is around 2, rathersmall changes in the interaction constants can affect thepredicted ground-state selection.

Viertio (1992) has made extensive Monte Carlo simu-lations of nuclear ordering in silver. For the [001] fieldalignment, she found a 1a→ 3 transition in intermediatefields, in agreement with Heinila and Oja (1993a). Invery small fields, B'0.1Bc , Viertio predicted that phase1a is unstable against another single-k state with the or-dering vector perpendicular to B.

b. Bi[110]

Heinila and Oja (1993a) have used the perturbationapproach to study the ground state for the [110] fieldalignment. The resulting phase diagram is shown in Fig.119.

It was found that the stable triple-k structures in thisfield direction have the general form

3b: 5m5m~ex1ey!/A2,

d15d1zez ,

d25d2zez ,

d35d3xex1d3yey ,

(138)

FIG. 119. Phase diagram for easy-plane type-I fcc antiferro-magnets, when the magnetic field is in the @110# direction, as afunction of m5B/Bc and d53D1 /(D12J1). Below the bro-ken line quantum effects tend to stabilize the triple-k structureof Figs. 113(b)–113(d) in an I5

12 system. From Heinila and

Oja (1993a).

Page 110: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

110 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

where m5B/Bc . A special case is the more symmetricconfiguration

3a: H d1z5d2z5@m~12m !#1/2,

d3x5d3y5~m21 !/A2(139)

predicted previously by a spin-wave calculation (Viertioand Oja, 1987).

Structure 3a is similar to the symmetric form of con-figuration 3 [see Eq. (137)] found for Bi@001# : it is afour-sublattice triple-k state in which spins in the twosublattices are parallel to the field for all B,Bc , viz.,

3a: 5I1 /I5~2m21 !~ex1ey!/A212@m~12m !#1/2ez,

I2 /I5~2m21 !~ex1ey!/A222@m~12m !#1/2ez ,

I3 /I5~1/A2 !~ex1ey!,

I45I3 .(140)

The sublattices 1–4 are defined in Fig. 117. In particular,configurations 3a and 3 [see Eq. (136)] are identical inthe isotropic limit d50, as they should be, because thecoupling between the spin directions and the crystallineaxes then vanishes. Structure 3a is illustrated in Fig.3(d). Note, however, that the field in the figure is inthe [011 ] rather than the [110] direction.

When B50, configuration 3b reduces to structure 2[see Eq. (135c)]. The transition in a low intermediatefield between 3b and 3a is discontinuous.

Finally, in the lower left corner of Fig. 119 the single-k structures 1z and 1y are stable. For these configura-tions the nonzero vectors dj are given by

1z: d35~pAF /A2 !~ex2ey!, (141a)

1y: d25pAFez , (141b)

where pAF5A12m2. In zero field 1z is identical with 1a[see Eq. (135a)] and 1y is the same as 1b.

The results of Fig. 119 are in fairly good agreementwith the spin-wave calculation by Viertio and Oja (1987,1989), but there is a clear disagreement with the earlierperturbation calculation of Lindgard (1988b, 1990).

In the d→0 limit the results for the anisotropic modelshould be identical with those found in the isotropiccase. The reason (Heinila and Oja, 1993a) why this isnot so in the diagrams of Figs. 116 and 119 is that thelimit d→0 is artificial in the sense that the conditions ofEq. (112) caused by the anisotropy of the interactionsare still required. As a result, part of the triple-k struc-tures are excluded from the analysis although single-kand double-k states are essentially included. For ex-ample, the triple-k phase of the isotropic model at m512 [see Fig. 113(c)] violates the constraints dj•kj50 of aneasy-plane system for all field alignments and thereforecannot be stable in the presence of an appreciable dipo-lar anisotropy.

When the anisotropy is weak the ground state ob-tained for the fully isotropic interaction is approximatelycorrect, and spin configurations of the isotropic modelreach into the (m ,d) phase diagram of the anisotropic

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

model for small d (Heinila and Oja, 1993a). Exactly howfar they extend depends on the relative magnitude ofquantum-mechanical fluctuations, which favor the triple-k phase of the isotropic model, and on the dipolar an-isotropy energy, which is a classical quantity. The phaseboundary in the (m ,d) plane depends, therefore, on thespin I .

According to estimates by Heinila and Oja (1993a), inthe region below the broken line in Fig. 119, quantumfluctuations stabilize the triple-k state of Figs. 113(b)–113(d). The region collapses towards the line d50 in the(m ,d) diagram with increasing I . For a given m , thequantity d along the phase boundary of this region isapproximately proportional to I21. The boundary ofthe quantum-fluctuation-stabilized phase depends onlyweakly on the field direction. It has been left out of Fig.116 for clarity.

Viertio (1990) has performed Monte Carlo calcula-tions to investigate nuclear magnetic ordering in silver.As in the simulations of Frisken and Miller (1988a), di-polar and RK interactions were included up to theeighth neighboring shell. The exchange constants weretaken from the band-structure calculations of Harmonet al. (1992). To distinguish the equilibrium structurefrom metastable states, Gibbs free energies were calcu-lated for stable and metastable states found in the simu-lations. This procedure permitted Viertio (1990) to con-clude that the equilibrium configurations are thoseshown in Fig. 120. In low fields, B,0.4Bc , she observedthat thermal fluctuations stabilize structure 1y of Eq.(141b). In high fields, B.0.4Bc , phase 3a correspondingto Eq. (139) was obtained. These results were in goodagreement with earlier, less extensive Monte Carlosimulations by Viertio and Oja (1989).

The Monte Carlo results can be compared with theperturbation calculation illustrated in Fig. 119 by notingthat the nearest-neighbor interaction assumed by Vi-ertio (1990) corresponds to d50.75. There is a discrep-ancy in low fields since the perturbation calculation pre-dicts a 1z configuration rather than 1y. In higher fieldsthe results are in qualitative agreement with each other.Using the perturbation approach to thermal fluctuations,

FIG. 120. Equilibrium spin configurations of silver in a mag-netic field oriented along the [110] crystalline direction asgiven by Monte Carlo simulations at T50.17 nK. In low fields,B,0.4Bc a single-k state (1y) is predicted; in high fields, B>0.4Bc , a triple-k state (3a) is suggested. The arrows indicate thedirections of nuclear magnetic moments 2ghIi . From Viertio,1990.

Page 111: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

111A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

one finds a similar discrepancy with Monte Carlo resultseven when the same long-range interactions are em-ployed (Heinila and Oja, 1994a). One can, therefore,conclude that, at least when thermal fluctuations areconsidered, the 1z configuration in low fields is an arti-fact caused by inadequacies of the perturbation ap-proach. One may expect that the same holds for quan-tum fluctuations, too, since a spin-wave calculation forsilver at T50 (Viertio and Oja, 1989) agrees well withthe Monte Carlo results. In spite of these and possiblysome other shortcomings, the perturbation approach re-mains useful, since it quite effectively provides the cor-rect overall picture.

c. Bi[111]

Fluctuation effects on antiferromagnetic structureswith only type-I order have not been investigated as ex-tensively for the [111] field alignment as for the [001]and [110] directions. The reason is that the latter twohigh-symmetry alignments were first studied in measure-ments using copper single crystals (Huiku et al., 1986;Jyrkkio, Huiku, Lounasmaa et al., 1988). When experi-ments were extended to the [111] direction (Annilaet al., 1992), it was found that in the high-field regionthere was no type-I order while in low and intermediatefields the (0 2

323) modulation was present. These results

made it clear that a study of fluctuation effects in type-Istructures alone would not be sufficient to determine themagnetic-spin configuration in this field direction. At thesame time, however, such a study might be useful inrevealing why type-I order is unstable only in this fielddirection. Therefore it is unfortunate that the [111] fieldalignment has not been investigated in as much detail asthe two other high-symmetry directions.

Ground-state selection in the [111] direction for cop-per and silver has been studied by spin-wave calcula-tions (Viertio and Oja, 1987, 1989) and by Monte Carlosimulations (Viertio and Oja, 1989; Viertio, 1992). Incopper the dipolar anisotropy is strong, whereas silver isclearly exchange dominated. The simulations predictedfor copper a complicated field evolution: single-k →triple-k (Pxy

b ) → single-k → triple-k (Pxyg ) → single-k,

while the spin-wave calculation yielded a simplerscheme: triple-k (Pxy

b ) → triple-k (Pxyg ). Here Pxy

b andPxy

g refer to the detailed form of the triple-k configura-tion, as defined in the paper by Viertio and Oja (1987).

In the case of silver, a single-k → triple-k transitionwas suggested by Monte Carlo simulations (Viertio andOja, 1989), as one would expect on the basis of the com-pletely isotropic model. The predicted triple-k state hasno special symmetry. In particular, it is not a configura-tion in which there are two sublattices oriented exactlyalong the field, as in the structure of Fig. 120 in highfields at T.0. Such a spin arrangement could not existin the [111] field direction owing to dipolar anisotropy.In the spin-wave calculations of Viertio and Oja (1987,1989), the ansatz of especially symmetric triple-k stateshas the adverse effect that single-k structures occupy toolarge a portion in the spin-wave-determined phase dia-grams when Bi@111# .

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

8. Comparison with experiments

One can compare the properties of the predicted spinstructures with neutron-diffraction measurements oncopper in the high-field region when B is along the [001]or the [110] crystalline axes. According to experimentaldata (Annila et al., 1992) for the [001] field alignment,the (1 0 0) intensity is strong when the field is perpen-dicular to the respective ordering vector [see Fig. 55(b)]but almost zero when the field is parallel to it [see Fig.55(a)]. This behavior can be understood, at least quali-tatively, if the stable structure in high fields is either 2 or1b. For structure 2 [see Eq. (135c)], the reflections per-pendicular to Bi@001# are

F2~1,0,0!5F2~0,1,0!} 12 ~ ud1u21ud2u2!5 1

2 ~12B2/Bc2!,

(142a)

while the parallel reflection is

F2~0,0,1!}ud3u250. (142b)

Exactly the same results are obtained for structure 1bwhen a domain average is taken over the two equivalentstaggered magnetization directions x and y in Eq.135(b). The neutron-diffraction data are, therefore, inagreement with the spin-wave calculations and MonteCarlo simulations of Viertio and Oja (1987, 1989) andwith the perturbation-theory results of Heinila and Oja(1993a) (d'2 for Cu) but not with the Monte Carloresults of Frisken and Miller (1988a)20 nor with the pre-vious perturbation calculations of Lindgard (1988a).21

Neutron-diffraction results for the [110] field direc-tions, illustrated in Fig. 57, are not so clear cut as thosefor the [001] directions. The (1 0 0) intensity is observedin high fields for both symmetrically inequivalent fieldalignments. The data show, however, that the (1 0 0)Bragg peak increases more steeply for Bi[1 01] [see Fig.57(c)] than for Bi[011] [Figs. 57(a,b)]. One should becautioned that the absolute intensities were not repro-ducible, as is shown by the different, although in prin-ciple equivalent, (1 0 0) Bragg intensities in Figs. 57(a)and 57(b). Apart from this problem the observed behav-ior is at least in qualitative agreement with the high-fieldconfiguration’s being the triple-k structure 3a, as waspredicted by the spin-wave calculations and MonteCarlo simulations of Viertio and Oja (1987, 1989) and byperturbation-theory results of Heinila and Oja (1993a).According to Eq. (139), the (1 0 0) and (0 1 0) neutron-diffraction intensities of 3a are

F2~1,0,0!5F2~0,1,0!} 12 ~ ud1u21ud2u2!

5~B/Bc!~12B/Bc!, (143a)

when Bi[110], and for the Bragg peak at (0 0 1) the in-tensity is

F2~0,0,1!}ud3u25~12B/Bc!2. (143b)

20Their phase AF4 shows behavior exactly opposite to thatactually observed.

21His high-field triple-k structure would show both perpen-dicular and longitudinal reflections.

Page 112: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

112 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

The (0 0 1) reflection should therefore be much weakerthan the (1 0 0) and (0 1 0) Bragg peaks in fields slightlybelow Bc , and its intensity should also increase moreslowly with decreasing B . The neutron-diffraction mea-surements are in disagreement with the perturbation cal-culations of Lindgard (1988b).22

In zero field, the three different type-I modulationsshould all have the same intensity. It is, however, obvi-ous from Fig. 56 that the direction of B during demag-netization affects the domain distribution and thereforemakes comparisons with theoretical calculations diffi-cult.

Theoretical phase diagrams for nuclear spins in silverhave been compared with experimental results in Sec.IX.C.

E. NMR response of type-I structures

The theory of the NMR response of easy-plane, type-Ifcc structures is important for an understanding ofnuclear ordering in copper and silver, and possibly inrhodium as well. The subject has been investigated theo-retically by Kumar, Kurkijarvi, and Oja (1986) and byHeinila and Oja (1995, 1996). Their work is based on thevast number of papers on antiferromagnetic resonancein electronic magnets that have been reviewed exten-sively by Foner (1963), Keffer (1966), and Akhiezeret al. (1968). However, there are several complicationsthat hamper direct application of this early work to thepresent case. Most of the NMR data on copper and sil-ver were measured in zero external field and hence theabsorption curves are broad. As the crystal structure iscubic in these metals, there is no single-spin anisotropyin the Hamiltonian, and the positions of the NMR linesare determined by spin-spin interactions and, within theordered state, by the spin structure. The situation is dif-ferent in most electronic systems in which antiferromag-netic resonance has been investigated: anisotropy causedby noncubic crystal symmetry and/or magnetoelastic ef-fects is important. In addition, in Cu and Ag one has todeal with the inherent frustration of the fcc lattice.

1. Equations of motion

In the mean-field theory, the spin dynamics is de-scribed by

d^Ii&dt

5g^Ii&3Bi , (144)

where Bi is the local field defined by Eq. (75). The equa-tions of motion for the amplitudes of the type-I modu-lations dj (j51,2,3) and for the magnetization m [seeEq. (111)] can be found by Fourier transforming Eq.(144). This yields

1g

dmdt

5m3B2BD (j51,2,3

~dj3 ej!~ ej•dj! (145)

22His high-field single-k structure would not give rise to the(1 0 0) reflection at all when Bi[011].

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

and

1g

ddj

dt5dj3~B2Bcm!2BD~m3 ej!~ ej•dj!

2BD~dk3 el!~ el•dl!2BD~dl3 ek!~ ek•dk!.

(146)

Indices (j ,k ,l) in Eq. (146) denote either (1,2,3),(2,3,1), or (3,1,2). The coefficient BD is the anisotropyfield for type-I order, defined by Eq. (120). When thedipole-dipole force is the only anisotropic contributionto spin-spin interactions, the tabulated (Cohen andKeffer, 1955) lattice sums yield for an fcc latticeBD'(m0/4p)rgI\36.501.

The anisotropy field BD changes the NMR responsein a profound way: the resonance peaks are shifted awayfrom the Larmor frequency. Isotropic spin-spin interac-tions alone, in a system with only one spin species, can-not cause any shifts from the Larmor position. The non-zero value of BD results from anisotropic spin-spininteractions like the dipolar force or an anisotropic ex-change interaction.

An important characteristic of copper and silver isthat BD.0. In other words, the anisotropy is such thatthe antiferromagnetic amplitudes dj are confined, inequilibrium, into planes perpendicular to the respectiveordering vectors kj , as is expressed by Eq. (112).

The similarity between these equations of motion andthose describing the dynamics of solid 3He and super-fluid 3He-A has been discussed by Kumar et al. (1986)and by Heinila and Oja (1996). If only one of the dj’s isnonzero, Eqs. (145) and (146) are almost identical withthose proposed for solid 3He (Cross and Fisher, 1985).The difference is that in 3He it is generally possible toneglect the terms proportional to the dipole-dipole en-ergy in Eq. (146) because they are small. This approxi-mation is not needed, however, when the equations ofmotion are solved to first order in the excitation field,since these terms disappear anyway. Equations (145)and (146) also have much in common with the Leggett(1974) equations of superfluid 3He-A if we identify hisl and d with our e1 and d1. Spin dynamics of 3He-Aresembles, however, a uniaxial antiferromagnet with aneasy axis along l and therefore the correspondingBD,0.

2. Resonances

The linear response can be calculated by consideringsmall deviations from the equilibrium configuration. Weassume that a field B(t)5B01B1(t) is applied to thesystem and expand the spin vectors in leading ordersof B1: m(t)5m(0)1m(1)(t)1O(B1

2) and similarly fordj(t). Our aim is to solve for the quantity m(1)(t), whichis of first order in B1(t). One obtains from Eq. (146)

1g

d

dt~ ej•dj!5~ ej3dj

~0 !!•~B12Bcm~1 !!1O~B12!. (147)

Combination of this result with Eq. (145) yields

Page 113: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

113A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

1g

d2m~1 !

dt2 5m~0 !3dB1

dt2B03

dm~1 !

dt2gBD

3 (j51,2,3

~dj~0 !3 ej!~ ej3dj

~0 !!•~B12Bcm~1 !!.

(148)

Introducing the Fourier transformation f(v)5(2p)21*2`

` f(t)exp(ivt)dt and noting thatM1(v)5rgI\m(1)(v) and M1

m(v)5(nxmn(v)B1n(v)/

m0 one finds, if vÞ0 (Heinila and Oja, 1996),

xH~v!/x05@Y~v!2Iv2/gBc#21Y~v!, (149)

where

Y~v!5gBD (j51,2,3

~ ej3dj~0 !!~ ej3dj

~0 !!

2iv (j51,2,3

~m~0 !3 ej!ej . (150)

The subscript H of xH(v) in Eq. (149) emphasizesthat the above calculation gives only the nondissipativepart of x(v), since the right-hand side of Eq. (149)makes up a Hermitian matrix. The dissipative part isabsent because there is no relaxation mechanism in theequations of motion.

Exactly at the resonant frequencies we must add themissing dissipation. The full dynamic susceptibility canbe written as

x~v!/x0512 (

a

La

12v/~Va2iGa!

112 (

a

La*

11v/~Va1iGa!1

L0

12iv/G0,

(151)

where Va are resonant frequencies and the Hermitianmatrices La are resonant amplitudes. In Eq. (149),xH(v) is the Hermitian part of the expression in thelimit Ga→10 when v Þ 0.

The third term in Eq. (151) describes absorption atzero frequency. Its physical meaning can be illustratedby considering the response to a small steplike change inthe external field. When the zero-frequency resonancehas, for example, the Lorentzian form, the system reactsvia a process in which the magnetization relaxes expo-nentially towards a new equilibrium value, and the re-laxation time is essentially the inverse of the linewidth atthe v50 resonance. The processes that contribute tozero-frequency absorption are those that involve achange in the spin-spin interaction energy but no changein the Zeeman energy. For NMR measurements of thev50 resonance, see Anderson (1962).

With zero-frequency resonance, and in general withlow-field experiments, one should look for resonances inx(v)/v rather than in x(v). The quantity x9(v)/v,where x9(v) is the imaginary part of a diagonal compo-nent of x(v), is often called the form function. For ex-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

ample, in the paramagnetic state in zero field, x9(v)/vpeaks at v50 if the temperature is well above Tc . Thispeak can be called the zero-frequency resonance.

Kramers-Kronig relations provide an equation be-tween resonant amplitudes La and the static susceptibil-ity,

x~0 !/x05(a

Re$La%, (152)

where x05m0rgI\/Bc [see Eq. (116)]. Equation (152)decomposes the static susceptibility matrix into contri-butions of various resonances Va , making it possible tocalculate the amplitude of the zero-frequency absorp-tion once the amplitudes of other resonances, as well asx(0), are known. Static susceptibility matrices of severaltype-I structures were discussed in Sec. XV.D.2.

As an important example, we consider a single-kstructure when the external field is not along the order-ing vector. The equation of the configuration is given byEq. (122). It turns out that x(v) can be expressed usingthe angle u between B0 and k as the only parameter.One obtains two resonant frequencies V1 and V2 givenby

V62 5 1

2 $v021g2B0

26A~v022g2B0

2!214g2B02v0

2cos2u%,(153)

where v025(BD /Bc)g2(p2Bc

22B02). This result was

found for antiferromagnetically ordered solid 3He byOsheroff et al. (1980) and later for metallic fcc nuclearmagnets (Kumar et al., 1986). A practical difference isthat v0 has a strong field dependence in nuclear magnetsbut can be taken as a constant in solid 3He.

The amplitudes of the two resonances of Eq. (153)were calculated by Heinila and Oja (1996). These au-thors also showed that the two amplitudes satisfy(aRe$La%5I . As the static susceptibility of the single-k state of Eq. (122) is simply x(0)5x0I (see Sec.XV.D.2), the spectral sum rule of Eq. (152) does notleave any room for zero-frequency absorption. The situ-ation is different, however, for a single-k structure atB50 or for a single-k state in a field when B0ik [see Eq.(117)]. Zero-frequency absorption is predicted.

As an application, one can study how well the reso-nance frequencies of a single-k state can explain the po-sitions of the antiferromagnetic NMR peaks that wereobserved by Hakonen and co-workers (Hakonen andYin, 1991; Hakonen, Yin, and Nummila, 1991). Such acomparison has been made by Heinila and Oja (1996).To account for the fact that a polycrystalline sample wasused in the NMR measurements, it was assumed that,among the three possibilities, the single-k structure forwhich the ordering vector maximizes (kj•B0)2 is stabi-lized in each field alignment. The assumption is sup-ported by the neutron-diffraction data of Tuoriniemi,Nummila, et al. (1995) as well as by most of the theoreti-cal predictions (Viertio and Oja, 1989; Viertio, 1990,1992; Heinila and Oja, 1993a, 1996). According to Eq.(153), the antiferromagnetic resonances form two bands,uvu<min$gB0 ,v0% and max$gB0 ,v0%<uvu<Ag2B0

21v02,

Page 114: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

114 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

when the direction of B0 is varied with respect to thecrystalline axes. Their calculated average frequencieshave been plotted in Fig. 71 assuming a sublattice polar-ization p50.7. There is rather good agreement. It seems,therefore, that the observed resonances can be inter-preted in terms of a single magnetically ordered phasewith a single-k structure. The same conclusion wasreached by Hakonen and co-workers.

3. Simulation of spin dynamics

Heinila and Oja (1995, 1996) have recently presenteda numerical calculation of the full NMR absorptioncurve of a classical spin system. Their technique appliesequally well to ordered and disordered systems. Themethod is similar to the procedures employed in the cal-culation of the dynamic structure factor (Wysin andBishop, 1990; Chen and Landau, 1994). Since themethod is, in principle, accurate for classical spins, it canbe used to assess the validity of the mean-field equation-of-motion analysis.

NMR response can be obtained from the autocorrela-tion matrix

C~ t !51N

^@M~ t01t !2^M&#@M~ t0!2^M&#&, (154)

where M5( iIi /I is the total spin. According to thelinear-response theory (Kubo and Tomita, 1954), C(t) isrelated to the dynamic susceptibility through

x~v!51

kBT@C~ t50 !1ivC~v!# , (155)

where

C~v!5E0

`

C~ t !eivtdt . (156)

The NMR response can be calculated numerically byusing Monte Carlo simulations to produce a set of spinstates. The time dependence of C(t) is then obtained bysolving the equations of motion [Eq. (144)]. For moredetails, see the papers by Heinila and Oja (1995, 1996).

As an example, the NMR response of a system withN5123 spins, modeling copper nuclei, was investigated.The simulations revealed a first-order transition to atype-I antiferromagnet at T5TN

MC. The spin structurewas a single-k state with spins along a crystalline axis,illustrated on the right-hand side of Fig. 4. The calcu-lated NMR absorption curves are presented in Fig. 121.The results show the presence of two antiferromagneticresonance peaks below TN

MC.The mean-field analysis presented in Sec. XV.E.2 pre-

dicts only a single antiferromagnetic resonance at a fi-nite frequency for this single-k state. The peak can beidentified with the high-f absorption signal, and it issimilar to conventional antiferromagnetic resonance.The peak at a low but finite f can be understood as anartificial zero-frequency resonance of the mean-field cal-culation. The failure of the theory is caused by a hiddensymmetry in the mean-field Hamiltonian [Eq. (74)],

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

which results in a continuously degenerate manifold oftype-I states, as was discussed in Sec. XV.D. This degen-eracy is lifted by thermal fluctuations—the kind of‘‘order-by-disorder’’ effect discussed by Villain et al.(1980)—and the single-k state with spins parallel to acrystalline axis is stabilized. In the dynamic susceptibil-ity, the order-by-disorder phenomenon is seen as a shiftof spectral intensity from f50 to a finite frequency.23

Therefore the anisotropy associated with the low-f peakis distinctly different from that of the upper peak.

As is implied by the theoretical curve at T50.09TNMC

(dotted line in Fig. 121), the width of the simulated reso-nance peaks tends to zero as T approaches zero. In areal system, however, narrowing of the peaks at lowtemperatures is limited by quantum fluctuations.

The theoretical line shapes can be compared with themeasured NMR data of Huiku et al. (1986) on Cu. Theexperimental spectra are shown in the inset of Fig. 121.The calculated paramagnetic line shape is in good agree-ment with experiments except for the too rapid falloff athigh frequencies. The measured spectrum displays tworesonances in the antiferromagnetic state, in agreementwith the simulations. The difference between the anti-ferromagnetic and paramagnetic states is, however,much smaller in the experiment than in the simulation.This could result from the coexistence of antiferromag-netic and paramagnetic domains during the measure-

23An analogous phenomenon is the disappearance of a gap inthe acoustic magnon dispersion relation at zero wave vector inthe frustrated electronic magnet Ca 3Fe 2Ge 3O 12 (Gukasovet al., 1988).

FIG. 121. Theoretically calculated zero-field NMR absorptionxxx9 (f) in the paramagnetic phase at T51.14TN

MC (solid line),and in the antiferromagnetic state at T50.86TN

MC (dashedline). Data for T50.09TN

MC (dotted line) have been divided by2, and the peaks on this curve are nearly d functions for thefrequency resolution used in the computations. These resultsare averages over different orientations of the crystal axes,namely, xxx9 (f)5Tr$Im$x(f)%%/3. The inset shows experimen-tal spectra of Huiku et al. (1986). More data are presented inFig. 32. All results are in SI units. From Heinila and Oja(1995).

Page 115: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

115A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

ment, since a suitable linear combination of the calcu-lated paramagnetic and antiferromagnetic spectraclosely reproduces the experimental line shape.

F. (0 23

23) order

Neutron-diffraction measurements of copper byAnnila et al. (1990) revealed, for the first time ever,ordering associated with the k=(p/a)(0, 2

3,23) modulation

vector in an fcc system. Theoretical studies were crucialin this discovery. First-principles electronic band-structure calculations by Lindgard, Wang, and Harmon(1986) had shown that ordering vectors of the formk=(0,h ,h) are almost degenerate in energy with type-Imodulation vectors in copper. Moreover, Lindgard(1988a) had predicted24 that this type of order would bestable in the intermediate-field region, exactly as wasfound in the experiments.

The knowledge of the ordering vector alone does not,however, yield the magnetic structure. The actual spinconfigurations with (0 2

323) order have been investigated

in extensive mean-field calculations by Viertio and Oja(1990a, 1990b, 1993). Their work has been remarkablysuccessful in explaining and predicting the properties ofspin structures with (0 2

323) order. Our discussion below

mostly draws from their papers.

1. Principal features of the (0 23

23) spin configurations

Analysis of the so-called permanent spin configura-tions, which were introduced in Sec. XV.B.4, has beenessential in the study of spin structures with (0 2

323) or-

der. The concept of permanent configurations, com-bined with the eigenvalue and eigenvector calculations,very effectively explains the principal features of the(0 2

323) order.

a. Easy-axis anisotropy

The (0 23

23) order is associated with easy-axis anisot-

ropy. This is seen, for example, by solving the eigen-value equation (103) in the J1J2D1 approximation, i.e.,when there are exchange interactions between the near-est and next-nearest neighbors and dipolar interactionsbetween nearest neighbors. For k5(p/a)(0, 2

3, 23) one

finds the following eigenvalues and eigenvectors:

l1523J123D1 , e15~1,0,0!, (157a)

l2523J116D1 , e25~1/A2 !~0,1,21 !, (157b)

l3523J123D1, e35~1/A2 !~0,1,1!. (157c)

Note that J2 does not enter into these equations. Thenearest-neighbor exchange interaction in copper is anti-ferromagnetic with J1 /kB'212 nK, while the dipolarcoupling D1 /kB5(m0/4p)\2g2r23/kB525 nK. There-fore l2 is clearly larger than l1 and l3, and the spin

24According to Heinila and Oja (1993a), the prediction wasbased, however, on an erroneous calculation, as explained inSec. XV.F.3.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

modulation associated with the (0 23

23) order is along the

eigenvector e25(1/A2)(0,1,21). The corresponding an-isotropy energy is, indeed, very large because the differ-ence of the eigenvalues is on the order of kBTN . The(0 2

323) order can therefore be classified as antiferromag-

netism with an easy axis.The important difference between anisotropy of this

sort and the structure found in traditional easy-axis an-tiferromagnets is that the anisotropy axis is dictated bythe ordering vector rather than by the lattice. Because ofthe cubic symmetry in copper, there are altogether 12different but symmetrically equivalent ordering vectors.These are the vectors in the star of k = (p/a)(0, 2

3, 23) (see

Fig. 54), namely, 6(p/a)(0, 23,6

23), 6(p/a)( 2

3,0,6 23),

and (p/a)( 23,6

23,0). As the axes associated with order-

ing vectors 6k are equivalent, there are altogether sixpossible easy axes. The presence of an external magneticfield however, breaks, the degeneracy between the vec-tors in the star.

b. Up-up-down structure

A simple example illustrates the general ideas. Let usconsider the single-k configuration

Ii /I5m

A2~0,21,1!1

d1

A2~0,21,1!cosFpa ~0, 2

3 , 23 !•riG ,

(158)

where d152 43 and m5 1

3; the structure is illustrated inFig. 3(c). It is a three-sublattice configuration in whichall spins have the saturation value u^Ii&u5I . The netmagnetization sums up to I/3 per spin. The antiferro-magnetic amplitude, d15(d1 /A2)(0,21,1), is along theeasy axis of the ordering vector k=(p/a)(0, 2

3,23). This

structure therefore seems a plausible candidate for theground state at T50 in a field aligned along the [01 1]direction. Exactly at B5Bc/3 the local fields at all spinsites are equal, and the structure is permanent. Then,according to Sec. XV.B.4, if l2(0, 2

3,23) is the largest ei-

genvalue, the up-up-down configuration minimizes theenergy and is indeed the ground state.

c. Superposition with type-I order

In fields lower than B5Bc/3 the system would like todecrease its magnetization from m5 1

3. Owing to theeasy-axis anisotropy of the (0 2

323) order, there is no way

to decrease m continuously while, at the same time, in-creasing the (0 2

323) component. Viertio and Oja (1990a)

suggested that the system then makes a linear combina-tion of the best and the second-best solutions: When thefield is lowered below Bc/3, the whole up-up-down pat-tern of Fig. 3(c) is reduced to decrease m , and the over-all length of the spin vectors is conserved by superposingthe up-up-down pattern with an up-down configuration,i.e., type-I modulation, in a perpendicular direction. Onepossible spin structure is then the one described in Fig.3(b).

When the field is reduced all the way to B50 there isonly type-I order left, as is shown by Fig. 3(a). Thewhole scenario can be described by using the equation

Page 116: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

116 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Ii /I5m

A2~0,1,21 !1

d1

A2~0,1,21 !cosFpa ~0,2

3 , 23 !•riG

1d2

A2~0,1,1!cosFpa ~1,0,0!•riG , (159)

where d1524m and d225129m2. One can prove that

this is the ground state under the following two assump-tions: (i) lmax5l2(0, 2

3, 23)5l2,3(1,0,0) and (ii) l2(0, 2

3, 23)

is larger than l2,3(1,0,0) by an infinitesimally smallamount.

One feature of Eq. (159), which is worth noting, isthat the amplitude of the (0 2

323) component is propor-

tional to the field. This particular type of antiferromag-netism is therefore induced by the field.

d. Construction of multiple-k structures

Viertio and Oja (1993) have derived equations thatcan be used to construct, in a general field direction,single-k and double-k spin configurations with only(0 2

323) order. The double-k structures can be written as

Ii /I5d1cos~k1•ri!1d2cos~k2•ri!1m, (160)

where k1 and k2 are vectors in the star of k5(p/a)( 2

3 , 23 ,0). The amplitudes of the antiferromag-

netic modulations d1 and d2 lie along the easy directionsof the respective ordering vectors k1 and k2 (see Sec.XV.F.1a). The magnetization is given by m5B/Bcwhere Bc5I@l2(0, 2

3,23)2l(0)]/\g .

At fcc lattice sites, the phase factor k•ri takes the val-ues 2p/3, 4p/3, and 2p , yielding cos(k•ri)52 1

2 , 2 12,

and 1, respectively. There are four sublattices with spins(d11d21m), (d12 1

2 d21m), (2 12 d11d21m), and

(2 12 d12 1

2 d21m). In zero field, m50 and it is clearlyimpossible to find amplitudes d1 and d2 such that thefour spin vectors would have equal lengths. However, atcertain values and directions of the external field, theinduced magnetization perfectly compensates for themismatch.

The analysis is performed at T50 for simplicity. Letus define m5mi1m' , where mi is the component ofmagnetization in the plane determined by d1 and d2. Thespins are saturated at Ii5I when the following condi-tions are satisfied:

d1•d250,

mi5214

~d11d2!,

9m i21m'

2 51. (161)

According to the first condition, the only possibledouble-k structure is of the type d1i(1,1,0) andd2i(1,21,0), with ordering vectors k15(p/a)( 2

3,223,0)

and k25(p/a)( 23, 2

3,0). The second condition of Eqs.(161) gives the correction mi , which is needed to makethe moments equal. There may be, perpendicular to the

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

plane spanned by d1 and d2, an additional component ofmagnetization m' , the value of which is determined bythe third condition.

In the case of a single-k structure, d250, the aboveequations reduce to mi52 1

4 d1 and 9m i21m'

2 51. It isclear that for a fixed direction of the external field, d1must have a component along B. Otherwise, antiferro-magnetic solutions are not found.

A similar analysis can be made for a superpositionstructure in which k1 is a vector in the star ofk5(p/a)( 2

3,23,0) and k2 is one of the three type-I order-

ing vectors. There are then four sublattices with spins(d11d21m), (d12d21m), (2 1

2 d11d21m), and

(2 12 d12d21m). At T50, Ii5I and the following con-

ditions must be satisfied:

d1•d250,

d2•m50,

mi5214

d1,

9m i21m'

2 1d2251. (162)

From the first condition and from the easy-axis anisot-ropy of type-I order, it follows that if d1i(1,1,0), thend2i(1,21,0) or i(0,0,1). The corresponding orderingvectors are k15(p/a)( 2

3,2 23,0) and k25(p/a)(0,0,1) in

the former case, and k25(p/a)(1,0,0) or (p/a)(0,1,0)in the latter. In a general field direction, in which umxu,umyu, and umzu are all nonequal and nonzero, it is thenimpossible to satisfy simultaneously the second equa-tion, and a double-k superposition structure does notexist. Once there is some symmetry in the field align-ment, for example, if mi[aab], a superposition solutioncan be found.

In the high-symmetry field directions, it is possible tofind more complex, permanent multiple-k structures thatsuperimpose the (0 2

323) and (1 0 0) orders. Viertio and

Oja (1993) have constructed, for the [100] and [011] fieldalignments, quadruple-k spin configurations that are si-multaneously modulated by one (0 2

323) ordering vector

and by the three type-I vectors.

2. Theoretical spin structure versus experiments on copper

Neutron-diffraction measurements on copper by An-nila et al. (1992), in which B was aligned along the vari-ous crystalline high-symmetry directions, revealed clearselection rules for the three type-I reflections and thetwelve cubic-symmetry-related (0 2

323) reflections (see

Fig. 59). These selection rules provided crucial tests forthe theoretically calculated spin configurations, as wediscussed in Sec. XV.D.8 for type-I configurations. Inthe case of the (0 2

323) order, the observed selection rules

provide even more tests of the spin structures. The ex-perimentally determined selection rules are in completeagreement with the theoretical predictions of Viertioand Oja (1990a, 1990b). Their equations for the actualspin configurations will be given below.

Page 117: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

117A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

a. Bi[011]

When the external field is aligned along the [011] crys-talline axis, there are three nonequivalent ordering vec-tors in the star of k25(p/a)( 2

3,23,0). The use of Eq. (161)

yields the following permanent single-k structures:

k15~p/a !~0, 23 ,2 2

3 !, d1524

3A2~0,1,1!, (163)

at B5Bc/3, and

k15~p/a !~ 23 ,0, 2

3 !, d1522

A6~1,0,21 !, (164)

at B5Bc /A3. Solutions similar to the latter can be ob-tained for the symmetrically equivalent ordering vectorsk5(p/a)( 2

3,0,2 23), (p/a)( 2

3,23,0), and (p/a)( 2

3,223,0).

For k5(p/a)(0, 23, 2

3), a permanent single-k solution doesnot exist.

Equation (161) also yields the double-k structure

k15~p/a !~ 23 , 2

3 ,0!, d152

A6~1,21,0!,

k25~p/a !~ 23 ,2 2

3 ,0!, d2522

A6~1,1,0! (165)

at B5Bc /A5.The above configurations are permanent only at dis-

crete values of B. In other fields it may be energeticallyadvantageous to combine the (0 2

323) modulation with a

suitable amount of type-I order, as was discussed in Sec.XV.F.1c. This is possible for the single-k structure of Eq.(163), as was demonstrated by Eq. (159) and in Fig. 3.The actual spin arrangement may be more complex,however. It is easy to see that one can superpose yetanother type-I modulation with the configuration givenby Eq. (159), namely, (d3,0,0)cos@(p/a)(0,1,0)•ri# , whered3 and d2 are now obtained from d2

21d325129m2. Fur-

thermore, it turns out that it is possible to add even athird type-I modulation, k5(p/a)(0,0,1), but the equa-tion for the spin configuration then becomes very com-plicated indeed.

The use of Eq. (162) shows that it is impossible tosuperimpose type-I order with the single-k configurationof Eq. (164) to construct a permanent double-k struc-ture.

The analytical solutions for spin arrangements by su-perposition are exact within the mean-field theory onlywhen lmax5l2(0, 2

3,23)5l2,3(1,0,0). Most of the degen-

eracy associated with the ground state is removed by theassumption that l2(0, 2

3,23) is larger than l2,3(1,0,0) by an

infinitesimally small amount. The ground state is thenthe one that maximizes the (0 2

323) order. This leaves,

however, some degeneracy in the spin system. A uniqueground state can be found by considering fluctuation ef-fects as was the case with pure type-I order. In addition,one has to take into account that the assumptionDl[l2(0, 2

3,23)2l2,3(1,0,0)→10 cannot be exactly valid

in a real system, and one has to accept a positive but

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

small Dl . One may consider this as an extra sourceof fluctuations to the ideal case Dl[l2(0, 2

3 , 23 )

2l2,3(1,0,0)→10. Monte Carlo simulations(Viertio and Oja, 1990a) have given evidence that thestable structure is a configuration in which thek15(p/a)(0, 2

3,223) ordering vector is superimposed on

k35(p/a)(0,1,0). The reason for illustrating in Fig. 3(b)the combination structure with the k25(p/a)(1,0,0) or-dering vector is that this configuration is easier to visu-alize.

Information on the stability of the various orderingvectors has been collected into Table VII. Comparisonwith the observed selection rules, which were summa-rized in Fig. 59, shows that there is complete agreementbetween theory and experiment: Only the predicted (Vi-ertio and Oja, 1990a, 1990b) ordering vectors were ob-served in measurements (Annila et al., 1992). In addi-tion, when Bi[011], the eight mutually equivalentvectors25 k56(p/a)( 2

3,0,6 23) and 6(p/a)( 2

3,623,0)

have their maximum intensities in higher fields than thek56(p/a)(0, 2

3,223) vector.

The observed field dependence of the (1 0 0) and the(0 2

323) Bragg reflections has been compared with calcu-

lations (Viertio and Oja, 1990a). The theoretical mag-netic structure factors uI(k)u2 are presented in Fig. 122 interms of a contour diagram in the B-T plane. Accordingto Eq. (51), uI(k)u2 is proportional to the intensity of theBragg peak, but experimental difficulties make it ex-ceedingly hard to obtain an accurate value for the pro-portionality constant. Measurements of relative intensi-ties are, however, very interesting as such because the

25Viertio and Oja associated these ordering vectors with thedouble-k configuration of Eq. (165) in their 1990a paper. Thestructure was then thought to be nonpermanent. Only later(1993) did the authors notice that the structure is, in fact, apermanent one. In their 1990a paper, the structure was called a4-k state rather than 2-k because 6k’s were counted sepa-rately. The single-k configuration of Eq. (164) was found onlyin the 1993 paper.

TABLE VII. Summary of theoretical selection rules for the (023

23) order. From Viertio and Oja (1990a, 1990b, 1993).

Alignment k(p/a) Stability region

Bi@100# 6(0, 23,6 2

3) not stable6( 2

3,0,6 23) 0 <B<Bc /A5

6( 23,6 2

3,0) 0 <B<Bc /A5Bi@011# 6(0, 2

3, 23) not stable

6(0, 23,2 2

3) 0<B<Bc/36( 2

3,0,6 23) Bc /A5, Bc /A3

6( 23,6 2

3,0) Bc /A5, Bc /A3Bi@111# 6(0, 2

3, 23) not stable

6(0, 23,2 2

3) 0<B<A3/19Bc

6( 23,0, 2

3) not stable6( 2

3,0,2 23) 0<B<A3/19Bc

6( 23, 2

3,0) not stable6( 2

3,2 23,0) 0<B<A3/19Bc

Page 118: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

118 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

contour diagram has a rich structure.There is clearly a good resemblance between the

theoretical and experimental (see Fig. 3) diagrams. Inhigh fields, only (1 0 0) order has been found. In inter-mediate fields, there is an abrupt transition between the(1 0 0) state and a structure with only (0 2

323) order.

When the field is lowered further the intensity of the(0 2

323) reflection decreases while the (1 0 0) peak in-

creases. Finally, at B50, only (1 0 0) order exists. Thetheoretical contour diagram reproduces well the two dif-ferent kinds of transitions between the (0 2

323) and (1 0 0)

structures: The transition is gradual in low fields andquite abrupt in high fields.

There are several matters, however, that one shouldpay attention to when making a comparison. First, thehorizontal axis in the experimental phase diagram is thewarmup time t rather than the absolute temperature.There has to be a monotonically increasing relationshipbetween t and T so that the topology of the two dia-grams should be the same. Another important point isthe way in which the theoretical diagram was con-structed. uI(k)u2 was obtained from a Monte Carlo simu-lation by heating the ground state in small temperaturesteps. The initial configuration was the combinationstructure of Eq. (159) in fields B,Bc/3, but with thek35(p/a)(0,1,0) ordering vector rather thank25(p/a)(1,0,0). In fields B.Bc/3, the initial configu-ration was the type-I structure 3a [see Eq. (139)].

The largest discrepancy between the experimentaland theoretical structure-factor diagrams occurs in fieldsB50.12–0.16 mT, where a gap develops upon warmingbetween the contours for the (0 2

323) and (1 0 0) Bragg

peaks. Viertio and Oja (1990a) suggested that this is dueto the fact that another structure, with 6( 2

3 0 6 23) and

6( 23 6 2

3 0) Bragg reflections, is stable in this field region(see the previous footnote). Subsequent experimental

FIG. 122. Contour diagrams of the magnetic structure factoruI(0, 2

3, 23)u2 (solid lines) and the averaged (1 0 0) structure fac-

tor 13@ uI(1,0,0)u21uI(0,1,0)u21uI(0,0,1)u2# (dashed lines) in the

temperature vs magnetic field plane for Bi[011 ]. The resultswere obtained in a simulated warmup during a Monte Carlorun. Compare with the experimental diagram shown in Fig. 3.From Viertio and Oja (1990a).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

results, summarized in Table VII, were in accord withthe suggestion. Apart from the double-k structure of Eq.(165), another plausible candidate for the stable struc-ture in the region of the gap is the single-k structure ofEq. (164).

b. Bi[100]

Permanent spin structures with (0 23

23) order have

been predicted for Bi@100# as well (Viertio and Oja,1990a, 1990b). In this field direction, too, it is possible toconstruct structures that superimpose the (0 2

323) and

(1 0 0) orders when l2(0, 23,

23)5l2,3(1,0,0). In these spin

configurations the (0 23

23) order is uniquely determined,

whereas the (1 0 0) order is continuously degenerate:various combinations of k=(p/a)(1,0,0), (p/a)(0,1,0),and (p/a)(0,0,1) modulations are possible. In the equa-tions below, the particular solutions that maximize theBragg intensities at positions (0 1 0) and (0 0 1) havebeen chosen (Viertio and Oja, 1993), since neutron-diffraction measurements (Annila et al., 1992) showedthat these reflections are much stronger than the one at(1 0 0) after demagnetization with Bi@100# , as illustratedin Fig. 56.

In fields 0,B,Bc/3 the permanent spin structurethat maximizes (0 2

323) order is the triple-k configuration

k15~p/a !~ 23 , 2

3 ,0!, d15d1

A2~1,21,0!,

k25~p/a !~2 23 , 2

3 ,0!, d25d2

A2~1,1,0!,

k35~p/a !~0,1,0!, d35d3~0,0,1!, (166)

d15d2522A2m ,

d356A129m2.

The structure is illustrated in Fig. 123(a). Viertio andOja (1990a) identified this spin configuration with theexperimentally observed low-field phase AF1 in Fig. 1.

FIG. 123. Spin structures in the AF1 and AF2 phases of cop-per when the external magnetic field is aligned along a crystal-lographic axis, here the @100# direction. The corresponding ex-perimental diagram is illustrated in Fig. 1. Only spins in onecrystallographic plane have been shown: (a) AF1 phase infields B,Bc/3 as given by Eq. (166); (b) AF2 inBc/3,B,Bc /A5 [see Eq. (167)]. From Viertio and Oja(1990a).

Page 119: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

119A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

In fields Bc/3,B,Bc /A5 one finds the quadruple-kstructure

k15~p/a !~ 23 , 2

3 ,0!, d15d1

A2~1,21,0!,

k25~p/a !~0,0,1!, d25d2

A2~1,1,0!,

k35~p/a !~0,1,0!, d35d3~0,0,1!,

k45~p/a !~1,0,0!, d45d4~0,0,1!, (167)

d1522A2m ,

d251

A2~m2A229m2!,

d356@ 12 m~A229m22m !#1/2,

d45d3 .

This configuration is illustrated in Fig. 123(b). The struc-ture was associated with the experimentally observedintermediate-field phase AF2.

In fields above Bc /A5, the ordered spin configurationAF3 should be the type-I structure with two orderingvectors [see Eq. (135c)], as was discussed in Sec.XV.D.7a. The stability regions of the various orderingvectors have been summarized in Table VII. Compari-son with observed selection rules of Fig. 59 again showscomplete agreement.

With the above identification of the three experimen-tally found phases in this field direction (see Fig. 1), thetransitions AF1↔AF2 and AF2↔AF3 are of first orderas observed. Even the absolute values for the transitionfields at T50 are in good agreement with experiment:For the exchange constants, which were used in theMonte Carlo simulations of Viertio and Oja (1990a),one finds Bc/350.13 mT and Bc /A550.18 mT, while themeasured values are 0.12 mT and 0.17 mT, respectively.

c. Bi[111]

Permanent spin configurations with (0 23

23) order have

been constructed for Bi@111# , too (Viertio and Oja,1990a, 1990b). In this case, there are no permanentdouble-k structures but a single-k solution was found atB5A3/19Bc . The ordering vector is perpendicular tothe field; it is either 6(p/a)(0, 2

3,223), 6(p/a)( 2

3,0,2 23),

or 6(p/a)( 23,2

23,0). These modulations are all equiva-

lent when the field is exactly along the [111] direction.The other six (0 2

323) ordering vectors do not yield per-

manent solutions. These theoretical selection rules arein agreement with the measurements (see Fig. 59).

In fields 0,B,A3/19Bc , the single-k structure can besuperimposed with (1 0 0) order, viz.,

k15~p/a !~0, 23 ,2 2

3 !, d15d1

A2~0,1,1!,

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

k25~p/a !~1,0,0!, d25d2

A2~0,1,21 !, (168)

d1528m/A6,

d256A1219m2/3.

d. Other field directions

Viertio and Oja (1993) have constructed all perma-nent spin configurations with one or two ( 2

323 0) ordering

vectors when the external magnetic field is applied in theplane By5Bz . The results of this calculation are illus-trated in Fig. 124. In the shaded region, there exists adouble-k configuration that can be obtained from Eq.(162) by a superposition of k15(p/a)(0, 2

3 ,2 23 ) and k2

5(p/a)(1,0,0) ordering vectors. Along the [011] and[100] field directions, the above mentioned quadruple-ksuperposition structures exist as well.

3. Stability of (0 23

23) modulation versus type-I order

There has been some controversy about the origin ofthe (0 2

323) order in copper. The question has been

whether this modulation is stabilized by fluctuation ef-fects beyond the mean-field theory, or whether the spin-spin interactions are such that the (0 2

323) order results

already from the mean-field theory. The latter possibil-ity would require that the indirect exchange interactionsdiffer by some 10 to 20% from their calculated and mea-sured values.

FIG. 124. Permanent spin configurations of copper with(0 2

323 ) order as functions of the external magnetic field in the

plane By5Bz . The structures exist along the heavy curveswith the particular ordering vector(s) indicated. The field val-ues (in units of Bc) for the high-symmetry directions are indi-cated. In the shaded region, (0 2

3 2 23) and (1 0 0) orders are

superimposed. The thin Bc curve shows the critical field forantiferromagnetic order. From Viertio and Oja (1993).

Page 120: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

120 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Although an ordering vector of type k5(p/a)(h ,h ,0) was a possibility in the mean-field cal-culations, which assumed the free-electron model for theRK interaction (Kjaldman and Kurkijarvi, 1979; Oja andKumar, 1984), this ordering became a serious alternativefor the modulation vector of copper only with the first-principles electronic-band-structure calculations of theRK interaction by Lindgard, Wang, and Harmon (1986).Their study predicted that the ordering vector should beof type I, as was thought previously, but also showedthat, if the strength of the RK interaction was somewhatreduced, the type-I ordering vector should become un-stable against the k5(p/a)(h ,h ,0) order rather thanthe k5(p/a)(h ,0,0) modulation, as was predicted bythe free-electron model (see Figs. 111 and 112).

In the first neutron-diffraction measurements on cop-per by Jyrkkio, Huiku, Lounasmaa et al. (1988), thetype-I Bragg reflection was found in the low- and high-field regions when Bi[01 1]; in intermediate fields theBragg intensity was almost zero. Soon afterwards, theo-retical calculations of Lindgard (1988a, 1988b) showedthat there are two different single-k states, 1z and 1y[see Eqs. (141)], in this field direction, with a transitionat B'0.3Bc . According to Lindgard, quantum fluctua-tions stabilize a k5(p/a)(h ,h ,0) phase in the field re-gion between the 1z and 1y states. This prediction ex-plained the low intensity of the type-I reflection inintermediate fields and was consistent with the (0 2

323)

Bragg reflection found in later neutron-diffraction ex-periments of Annila et al. (1990). In a more recent pa-per, Lindgard (1990) confirmed that the (0 2

323) order is,

indeed, favored by quantum fluctuations.Different results were obtained, however, in the re-

cent work by Heinila and Oja (1993a). They noticed thatLindgard’s calculation of the ground-state energies fortype-I structures was partly incorrect. According to Hei-nila and Oja, perturbation theory predicts the type-Istructures 3b and 3a for copper (d52) when Bi[011 ](see Fig. 119) rather than 1z and 1y. Furthermore, theseauthors found that quantum fluctuations favor type-I or-der rather than the (0 2

323) modulation. At least in the

important case of B5Bc/3, for which measurementsshow that the (0 2

323) Bragg peak has its maximum inten-

sity, the perturbation calculation of Heinila and Oja pre-dicts, in the case of complete degeneracy l2,3(1,0,0)5l2(0, 2

3 , 23 ), that quantum fluctuations favor the 3a

structure rather than the up-up-down configuration of(0 2

323) order. This work therefore gives support to the

idea advocated earlier by Viertio and Oja (1990a, 1990b,1993), who explained the observed stability of the(0 2

323) order by starting from the assumption that the

exchange interactions are such that the eigenvalue for k5(p/a)(0, 2

3, 23) is, in fact, larger than that for a type-I

ordering vector.The problem that remains is that the eigenvalue for

the (0 23

23) spin configuration is 10% lower than that for

type-I order according to first-principles band-structurecalculations (Lindgard et al., 1986; Oja et al., 1989).Since the 10% difference in the two eigenvalues iswithin the estimated uncertainty of the calculated ex-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

change parameters, there is no serious discrepancy. Nev-ertheless, refined computations of exchange interactionsin copper would be of value.

According to band-structure calculations made so far,several exchange parameters seem to be important indeciding the relative stability of the (0 2

323) and (1 0 0)

orders, since their energies are so close to each other. Itwould be particularly important to narrow down errormargins for the isotropic interactions between the near-est and the third-nearest neighbors. Anisotropic ex-change parameters between nearest neighbors are sig-nificant, too, as well as other isotropic interactions evenup to the sixth-nearest neighbors.

4. Superposition structure or a mixed stateof (0 2

323) and (1 0 0) domains

One of the main features of the neutron-diffractiondata on copper is that in a wide-field region, approxi-mately from B50.02 to 0.06 mT, both (0 2

323) and (1 0 0)

orders are present simultaneously. There are two inter-pretations of this fact. First, there can be superpositionof (0 2

323) and (1 0 0) orders in a single magnetic domain,

as was discussed in Sec. XV.F.1c. Without theoreticalsupport, such a complicated spin configuration might ap-pear unlikely. Second, the ordered spin system can be amixture of two different kinds of domains, namely, onewith (0 2

323) order and the other with (1 0 0) modulation.

It was stressed by the experimentalists (Annila et al.,1990, 1992) that measured data cannot distinguish be-tween these two possibilities.

Assuming tentatively the second alternative, Annilaet al. interpreted their data as a first-order transition be-tween (0 2

323) and (1 0 0) orders at B'0.06 mT, with

large hysteresis effects. The (0 23

23) Bragg reflection in

fields below B50.06 mT would then result from meta-stable domains. The relaxation time to thermodynamicequilibrium would have to be longer than the warmuptime caused by spin-lattice relaxation.

If the possibility of a superposition structure is ac-cepted, as is suggested by theoretical calculations (Vier-tio and Oja, 1990a, 1990b, 1993), there is no need toassume large hysteresis effects. The simultaneous pres-ence of (0 2

323) and (1 0 0) orders is thereby explained, as

shown by the good agreement between the theoreticaland the experimental structure-factor diagrams (Figs. 3and 122). Annila et al. (1990, 1992) conclude that a su-perposition structure is possible particularly in the fieldregion from B50.02 to 0.05 mT, where the temporalbehavior of the (0 2

323) and (1 0 0) Bragg reflections is

similar, as shown by Fig. 46(b). Although this is a com-pelling argument, it should be noted that, according toMonte Carlo simulations (Fig. 122), the T dependenceof the two Bragg reflections is different. Therefore thesuperposition structure could, in principle, be stable alsoin fields higher than 0.05 mT.

In very low fields, below 0.02 mT, the intensity of the(0 2

323) Bragg reflection rapidly decreases, as is shown by

curve (d) in Fig. 45. Depending on the nature of the(0 2

323) order in fields above 0.02 mT, this behavior sug-

Page 121: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

121A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

gests that the superposition structure becomes unstableor, alternatively, that the (0 2

323) domains disappear.

There are some measurements that could be useful indeciding between the superposition structure or a mixedstate. One possibility relies on the fact that Eqs. (162)show that it is impossible to construct, in a general fielddirection Bi[a b g] where uau, ubu, and ugu are all differ-ent and nonzero, a permanent double-k spin structurethat would superimpose the (0 2

323) and (1 0 0) orders. It

is likely that more complicated, permanent multiple-kspin configurations do not exist either. On the otherhand, a superposition structure is predicted for the cubichigh-symmetry field directions. One would then expectthat, for a suitable general field alignment, the behaviorof the (0 2

323) and (1 0 0) reflections in low intermediate

fields would be different from their behavior in high-symmetry directions.

Another possibility is to perform NMR measurementson a single-crystal specimen. A careful analysis of theNMR line shape might help in deciding between the twopossibilities (Heinila and Oja, 1995).

5. Related electronic magnets

The simultaneous presence of two different kinds ofordering vectors is a rare occurrence. Moreover, a pe-riod of three is rather unusual in ordered spin structures.The purpose of this section is to draw attention to someelectronic magnets sharing these properties in the hopethat this will result in a better understanding of thesesystems.

In TbMn 2, which crystallizes in the C15 cubic Lavesphase, the Mn atoms are distributed on the vertices ofregular tetrahedra, which, in turn, are packed in a dia-mond arrangement sharing vertices. A high degree offrustration arises for the Mn antiferromagnetism, lead-ing to complicated magnetic structures (Ballou et al.,1988). Neutron-diffraction measurements have revealedthat the low-temperature magnetic phase of TbMn 2 ismetastable and poised between magnetic configurationswith propagation vectors ( 2

323 0) and ( 1

212

12), respectively

(Ballou et al., 1992). Experiments have also shown thatthese reflections depend on the external magnetic fieldin an interesting way.

A period-of-three modulation has also been foundfor USb 0.9Te 0.1 in which the U atoms occupy fcc sites.Neutron-diffraction experiments have shown that,in the low-temperature phase, the ordering vectors6(p/a)( 2

3,0,0), 6(p/a)(0, 23,0), and 6(p/a)(0,0, 2

3), aswell as a ferromagnetic component, are present simulta-neously (Rossat-Mignod et al., 1979).

Tripling of the magnetic unit cell along one crystallineaxis has been observed in MnSe 2 (Hastings et al., 1959).Mn atoms occupy fcc sites in this compound having apyrite structure. The antiferromagnetic state displays theordering vectors (p/a)( 1

3,0,1) and (p/a)(0,1,0). Thetype-I component in this case is the third harmonic ofthe fundamental ordering vector (p/a)( 1

3,0,1). The roleof anisotropic interactions in MnSe 2 has been discussedtheoretically (Heinila and Oja, 1994c).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Finally, we comment on the up-up-down configurationpredicted for copper in a field B5Bc/3 when Bi[011 ][see Fig. 3(c)]. A similar configuration with the propaga-tion vector along a cubic axis has been found in calcula-tions on Ising systems for certain nearest and next-nearest neighbor interactions (Binder et al., 1981). Thenature of this structure, however, is different from thatfor vector spins.

Experimentally, an up-up-down spin configuration hasbeen observed in EuSe, which is an fcc system (Griessenet al., 1971, and references therein). In contrast to cop-per, magnetoelastic energies play a significant role in thebehavior of this compound.

G. High-field phase of copper when Bi[111]

Neutron-diffraction measurements of Annila et al.(1992) on copper showed unexpectedly that in fields B50.17 –0.25 mT, applied along the [111] crystalline di-rection, no antiferromagnetic (1 0 0) Bragg peak was ob-served (see Fig. 58), although it was found for the [100]and [011] field directions. Neither was there any (0 2

323)

structure present. Several other ordering vectors weresearched for but with negative results, as was describedin Sec. VII.J. A number of theoretical calculations havebeen performed to explain the absence of the (1 0 0)order and to predict the actual spin configuration.

1. A (h k l) structure

Assuming that the soft-mode theory, presented in Sec.XV.B.6, may be used to describe the transition from thepolarized phase to the ordered state, Oja and Viertio(1992, 1993) have predicted the form of the orderingvector for the as-yet-unknown Bi(1,1,1) phase. First,lmax must be obtained for a vector other than k=(p/a)(1,0,0). Otherwise the transition would alwaystake place at a type-I vector for all field alignments,since l2,3(p/a ,0,0) is degenerate. Clearly, lmax must bea unique eigenvalue. The corresponding eigenvector eshould not be perpendicular to any of the three [100]directions so that type-I order could be stable in thesefield alignments. Therefore ex Þ 0, ey Þ 0, and ez Þ 0. Forthe same reason, e should not be perpendicular to any ofthe six [110] directions. Thus uexuÞueyu, uexuÞuezu, andueyuÞuezu. The eigenvector corresponding to the un-known ordering vector Q is thus of most general form.Using group-theoretical arguments or direct calcula-tions, one can show that Q itself has to be of the mostgeneral form, namely, Q5(h ,k ,l), with uhu, uku, and uluall unequal and nonzero.

a. Model Hamiltonian

A Q5(h ,k ,l) type of ordering vector may seem toocomplicated for an fcc system. Viertio and Oja (1993),however, have performed extensive numerical calcula-tions to demonstrate that such an ordering vector is, atleast in principle, a real possibility. As the first step,theoretical spin-spin interactions in copper were alteredto produce a model that would favor ordering at a Q=

Page 122: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

122 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

(h ,k ,l) type vector. Modifications of the exchange con-stants were arbitrary except for the fact that changeswere kept within limits that are reasonable in view ofthe state-of-the-art band-structure calculations of ex-change interactions (Lindgard et al., 1986; Oja et al.,1989). It turned out to be possible to construct, usinginteractions between spins up to the ninth-nearest-neighboring shell, a Hamiltonian for which the followingtwo inequalities hold:

lmax5lmax~23 , 1

2 , 16 !*l2~0,2

3 , 23 !*l2,3~1,0,0!. (169)

Here, the set of k-vectors has been restricted to thosecompatible with a system of 123/25864 spins in an fcclattice, which is the reason that Q assumes the rationalvalue Q=(p/a)( 2

3, 12,

16). The inequalities for l2(0, 2

3,23),

which is the relevant eigenvalue for the (0 23

23) order, are

needed to explain the interplay of the (0 23

23) and (1 0 0)

modulations in intermediate fields.The interaction constants that produce this situation

were tabulated in the paper of Viertio and Oja (1993).When they were compared with the calculated exchangeconstants of Oja, Wang, and Harmon (1989), it wasfound that the largest absolute change occurred in thenearest-neighbor interaction, which was enhanced by18%. The modified values are thus reasonable in thesense that they fall within the error bars of the first-principles calculations and within the uncertainty of thevarious experiments probing exchange constants.

For the modified interactions, the soft-mode transitiontakes place to a Q=(p/a)( 2

3 , 12 , 1

6 ) state in fields alignedclose to the [111] crystalline direction, as shown by Fig.125. Whereas in the [011] and [100] field alignments,

FIG. 125. Critical fields Bc for antiferromagnetic structures incopper, with ordering vectors in the star of Q=(p/a)( 2

3, 12, 1

6), asa function of the external-field direction in the planeBy5Bz . The six ‘‘finger’’-shaped curves represent criticalfields for the various cubic-symmetry related vectors. The bro-ken lines with labels (1,0,0) and (0, 2

3, 23) show Bc’s for the cor-

responding ordering vectors. The scale is Bcmax50.37 mT. The

resolution has been enhanced greatly as shown by the valueson the horizontal axis. From Oja and Viertio (1993).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Bc for a type-I structure is higher than for a Q= (p/a)( 2

3,12, 1

6) configuration. An additional complication arisesfrom the fact that Bc(0, 2

3,23) is higher than Bc(1,0,0).

However, a suitable 2° misalignment of the field awayfrom the plane By5Bz stabilizes type-I modulation overthe (0 2

323) order in the [011] and [100] field alignments,

as observed.

b. Interplay of the (h k l), (0 23

23), and (1 0 0) modulations

Since the soft-mode theory applies only along thephase boundary B5Bc , several questions could not beanswered. To obtain more information on the actualspin configurations and the magnetic phase diagram ofcopper nuclei, Viertio and Oja (1993) performed nu-merical mean-field calculations in the antiferromagneti-cally ordered region, for different alignments andstrengths of the external magnetic field.

The stable spin configurations were determined by nu-merical iteration of the mean-field equations. The pro-cesses started from various initial configurations made ofseveral combinations of the three modulations (2

312

16),

(0 23

23), and (1 0 0). The iterations were continued until a

self-consistent solution was found. At every step, one ofthe 864 spins was randomly chosen, and a new directionand value was assigned to it according to Eqs. (75)–(77).The temperature was fixed at 140 nK=0.83 TN

MF whichcorresponds to the sublattice polarization u^Ii&u/I50.61.

At each field point, 7–9 different initial configurationswere tried. Typically 5–7 inequivalent final states werefound, and the stable structure was extracted from themetastable ones by comparing their Gibbs free energies.The phase diagram as a function of the field was thenconstructed in the plane By5Bz . The result turned outto be very complex. Altogether 17 ordered phases werefound. The magnetic phase diagram deduced from thesecalculations is presented in Fig. 126.

The ordering vectors of the different phases havebeen sorted out in Table VIII. The phases are labeled byA, B, and/or C to show that the structure contains( 2

312

16), (0 2

323), or (1 0 0) orders, respectively. It can be

seen that most of the phases are complicated superposi-tion structures of two different types of ordering vectors.For information on the modulation amplitudes, we referthe reader to the original publication (Viertio and Oja,1993).

In high fields, close to the [111] direction, phasesA1–A4 contain only ( 2

312

16) type modulations. This is

different from the [100] and [011] field directions, inwhich the (1 0 0) order, combined with the ( 2

312

16) modu-

lation, appears up to Bc . The results are in agreementwith the neutron-diffraction measurements of Annilaet al. (1992).

The simulations revealed some interesting nonlineareffects that were not anticipated on the basis of the lin-ear soft-mode analysis. In particular, exactly in the [100]and [011] field directions, soft-mode analysis predictedthat the (0 2

323) order would become stable at higher

fields than the (1 0 0) order. The simulations showed,

Page 123: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

123A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

however, that the stable structure combines (1 0 0) and(2

312

16) modulations but not the (0 2

323) component.

Studies of metastable states also produced some inter-esting and possibly important features. A pure ( 2

312

16)

state is not even metastable in high fields whenBi@100# or Bi@011# . In contrast, the pure high-field(1 0 0) structure was found to be metastable in the [100]and [011] directions but unstable in the [111] direction.

There is a discrepancy between the theoretical andexperimental phase diagrams of Figs. 126 and 61 in smalland zero fields. Although the fundamental orderingvectors in the (ABC) phase are (p/a)( 1

6,23, 1

2) and(p/a)(0,1,0), the structure also contains somek5(p/a)( 2

3,23,0) order at zero field, even though the

measured intensity of this Bragg peak vanished atB50. There are two ways to resolve this discrepancy. Itis possible that the first-order transitions (BC)1→(ABC) and (BC)3→ (ABC) (see Fig. 126) do not takeplace during the experimental time scale, so that phases(BC) 1 and (BC) 3, which are formed when B is loweredto zero during adiabatic demagnetization, are meta-stable at B50. The structures (BC) 1 and (BC) 3 in zerofield become pure (1 0 0) states, which are indeed meta-stable according to the simulations. Another possibilityis that the (p/a)( 2

3, 23,0) component is just the third

harmonic of the two fundamental modulations(p/a)( 1

6,23, 1

2) and (p/a)(0,1,0). The appearance of the(2

323 0) order at B50 could therefore be due solely to the

particular value k=(p/a)( 16,

23, 1

2), yielding lmax in thesimulations. If the maximum eigenvalue were located atanother k5(h ,k ,l), the third harmonic component

FIG. 126. Magnetic phase diagram of nuclear spins in copperat T50.83TN

MF in the plane By5Bz , as given by numericalmean-field simulations of Viertio and Oja (1993). The diagramconsists of 17 phases, which are labeled by A, B, and/or C toshow that the structure contains ( 2

312

16), (0 2

323), and/or (1 0 0)

order, respectively. For clarity, similar phases have beengrouped into larger units using heavy lines for their bound-aries; other borders are indicated with dashed curves. The par-ticular ordering vectors in the stars of the ( 2

312

16), (0 2

323), and

(1 0 0) modulations are listed in Table VIII.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

TABLE VIII. Ordering vectors for the theoretical spin struc-tures illustrated in Fig. 126. Most phases can have differentdomains, but only one is listed; others are obtained by theappropriate symmetry transformations, for instance, by ex-changing the y and z components of the ordering vectors.From Viertio and Oja (1993).

Phase Ordering vectors k(p/a)

A1 ( 16, 2

3, 12) & ( 1

2, 23, 1

6)

A2 ( 23, 1

2, 16)

A3 ( 23, 1

2, 16) & ( 2

3, 16, 1

2)

A4 ( 23, 1

2,2 16) & ( 1

6, 23, 1

2)

B1 ( 23, 2

3, 0)

B2 ( 23, 2

3, 0) & ( 23,2 2

3, 0)

C (1,0,0) & (0,1,0) & (0,0,1)

(ABC) ( 16, 2

3, 12) & ( 2

3, 23, 0) & (0,1,0) or

( 12, 2

3,2 16) & (0, 2

3,2 23) & (0,1,0) or

( 23,2 1

6, 12) & ( 2

3,2 23, 0) & (1,0,0)

(AB) ( 23, 2

3, 0) & ( 16, 2

3, 12)

(AC)1 5~

12, 2

3, 216!

~12, 2

23, 1

6!

~12, 2

16, 2

3!

~12, 1

6, 223!6 & (1,0,0)

(AC)2 5~

12, 2

3, 16!

~12, 2

23, 2

16!

~12, 1

6, 23!

~12, 2

16, 2

23!6 & (1,0,0) when Bx<By

(AC)2 5~

23. 1

6, 12!

~223, 2

16, 1

2!

~16, 2

3, 12!

~216, 2

23, 1

2!6 & (0,0,1) when Bx>By

(AC)3 5~

23. 2

16, 1

2!

~223, 1

6, 12!

~16. 2

3, 12!

~21

6, 223, 1

2!6 & (0,0,1)

(AC)4 5~

16, 2

23, 1

2!

~216, 2

3, 12!

~16, 2

3, 12!

~216, 2

23, 1

2!6 & (0,0,1)

(AC)5 ( 16, 2 2

3, 12) & (1,0,0)

(BC)1 ( 23, 2

3, 0) & ( 23, 2 2

3, 0) & (0,1,0)

(BC)2 ( 23, 2

3, 0) & (1,0,0) & (0,1,0) & (0,0,1)

(BC)3 (0, 23,− 2

3) & (1,0,0)

Page 124: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

124 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

would not be (p/a)( 23,

23,0). Actually, the (ABC) phase

would then be of type (AC).As for quantitive features, the agreement between the

phase diagram of Fig. 126 and the neutron-diffractionmeasurements could be somewhat better. For example,the configurations A 1–A 4 with no (1 0 0) order shouldextend to lower fields in the [111] direction. There are,however, possible explanations of this discrepancy (Ojaand Viertio, 1993).

In low intermediate fields, the phase diagram of Fig.126 reproduces the interplay of the (0 2

323) and (1 0 0)

modulations consistently with the experiments of Annilaet al. (1990, 1992) and the earlier theoretical calculationsof Viertio and Oja (1990a, 1990b).

All in all, these mean-field results can model the verycomplicated behavior of copper nuclei in the wholeBy5Bz plane within a single theoretical framework us-ing fixed values of interaction parameters.

c. Criticism

It has been emphasized by Lindgard (1992) that itwould be more satisfactory if the [111] phase of coppercould be explained using a simpler set of exchange pa-rameters than those employed in the simulations ofViertio and Oja (1993). However, the first-principles cal-culations of the spin-spin interactions have shown thatthe long range of the isotropic Ruderman-Kittel interac-tion, as well as anisotropic exchange forces, can be im-portant in the selection of the ground state. In fact, theuse of a linear theory, together with the constraints ob-tained from experiments, dictates the most symmetricform Q5(h ,k ,l) for the ordering vector. A simpler typeof order can emerge only as a result of fluctuations,which are neglected in the linear soft-mode calculationand in the mean-field theory.

Fluctuations might indeed play a decisive role in thisproblem. For the interactions used in the mean-fieldcalculations of Viertio and Oja (1993), the energiesof the three different types of modulations are veryclose to each other: The relevant eigenvalues arelmax(

23 , 1

2 , 16 )/kB=135.57 nK, l2(0, 2

3 , 23 )/kB=135.31 nK,

and l2,3(1,0,0)/kB=135.16 nK. Monte Carlo simulationscould, of course, take into account thermal fluctuationsof classical spins. However, the problem in such simula-tions would be the presence of several metastable states,and determining the stable structure would be computa-tionally a difficult task. The simulations using mean-fieldtheory did not suffer from this problem since calculationof the Gibbs free energy was straightforward.

It should also be noted that all theoretical work doneso far has considered only the conventional, long-rangeantiferromagnetic order.

In the experiments, on the other hand, one of the im-portant questions is whether the observed behavior re-lates to equilibrium situations. It is difficult, if not impos-sible, to fully assure oneself that this is indeed the case.Even for electronic magnets, with several decadesshorter time scales, nonequilibrium behavior has beenobserved in geometrically frustrated systems (Ballouet al., 1992).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

2. Other possibilities

The high-field spin structure of copper for Bi[111] hasbeen studied by Lindgard (1992). He, too, employed thesoft-mode theory. A detailed comparison between hisresults and the calculations of Oja and Viertio (1993) isdifficult because Lindgard reported in his brief paperonly the final numerical results.

Lindgard (1992) found that the wave vector of thespin-wave excitation, which softens in the paramag-netic state, is (i) the type-I vector (p/a)(1,0,0),(p/a)(0,1,0), or (p/a)(0,0,1) when Bi@001# or

Bi@110# , and (ii) k5(t1d ,t2d ,0) with t;( 23 )(p/a) and

d;0.05(p/a) when Bi@111# . These results conflict, how-ever, with some analytical properties of the soft-modetransition discussed in Sec. XV.B.6, and they cannot bereproduced from this theory (for further details, see Ojaand Viertio, 1993). Although the results of Lindgardwere promising in the sense that they were consistentwith neutron-diffraction data and provided an interest-ing prediction to be tested by future experiments, itseems that his results are erroneous.

Testing the prediction of a Q5(h ,k ,l) type orderingvector is very difficult, since it is not practical to scanthrough the whole reciprocal space. For experimentalistsit would therefore be desirable for the order to show upin a more accessible, high-symmetry direction of the kspace. This, together with the critical remarks on the(h k l) order presented in the previous section, has beenthe motivation for other theoretical suggestions.

One proposal (Oja and Viertio, 1992) is based on theidea that the lock-in mechanism that stabilizes the(0 2

323) order at a commensurate k vector in low and in-

termediate fields is not operational in high fields. This isplausible in view of some observations on electronicmagnets, namely, that a commensurate order → incom-mensurate order → disorder sequence is often foundwith increasing T , when the low-T ordering vector is notat the zone boundary (see, for example, Rossat-Mignodet al., 1979). When copper nuclei enter the high-field[111] phase, they are necessarily near the ordering tem-perature. By examining the energy eigenvalues, Oja andViertio concluded that incommensuration is most likelyto occur for a k vector of the form k=(p/a)(d, 2

3+e, 23 +e).

In fact, a local, but not the global, maximum of ln(k) isobtained at k = (p/a)(0, 0.55, 0.55), making the @0hh#direction most promising, as is illustrated by Fig. 112.Neutron-diffraction measurements testing this possibil-ity have not as yet been performed (see Sec. VII.J).

It has also been suggested that the high-field [111]structure of copper could be a spiral configuration withthe ordering vector along the field direction (Oja andViertio, 1992; Viertio and Oja, 1993). Such a proposalfaces difficulties in low fields and requires a near degen-eracy of three completely different types of orderingvectors. In addition, the calculated exchange interac-tions would have to be off by tens of percent.

Siemensmeyer and Steiner (1992) have suggested thatmagnetoelastic energy due to quadrupolar moments of

Page 125: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

125A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

copper nuclei is important in the selection of the stablespin structure in the high-field region.

H. Ferromagnetic ordering at T<0

Ferromagnetic ordering at negative temperatures fora spin system interacting through the dipolar and ex-change interactions, as in silver (see Hakonen, Num-mila, Vuorinen, and Lounasmaa, 1992), has been ana-lyzed theoretically by Viertio and Oja (1992). Theirwork follows the earlier treatments of nuclear-spin sys-tems of insulators in a rotating coordinate frame(Abragam and Goldman, 1982). There are, however,some important differences between the two cases ow-ing to the fact that only the truncated part of the dipolarinteraction is important for insulators.

At T,0, the system orders into a state that corre-sponds to the minimum eigenvalue ln(k), provided thatthis state is permanent. The minimum value can be ob-tained in the limit k→0 (see, for example, Kumar et al.,1985),

lmin[l1~k→0 !5~2 23 1R !m0\2g2r , (170a)

e1~k→0 !5k. (170b)

The limiting values are to be taken at an infinitesimallysmall but nonzero k. Exactly at k50, the eigenvaluelm(0)5(L2Dm1R)m0\2g2r depends on the shape ofthe sample.

For an infinite plate perpendicular to the z direction,Dx5Dy50 and Dz51, so that l1(k→0) coincides withlz(0). Therefore a k50 state of uniform magnetizationalong the z direction is stable.26

For all other sample shapes l1(k) is discontinuous atk50, and the maximum eigenvalue is obtained for avanishingly small but nonzero uku. Physically, this meansthat the sample breaks into domains that can be de-scribed by a quasipermanent spin structure, which is asuperposition of modulations with small k-vectors(Abragam and Goldman, 1982). In contrast to the situ-ation in dielectrics, the directions of k and ^I(k)& in ametal are not restricted by the external magnetic field(Viertio and Oja, 1992). This results in a wide variety ofstable structures.

1. Domain configurations

Let us consider the case of two types of domains withspins ^Ii&5IA in domains of type A and ^Ii&5IB in do-mains of type B. Three possible two-domain configura-

26It is interesting that for a purely dipolar system in an fcclattice lmax is also obtained at k→0: l2,3(k→0)5m0\2g2r/3,with e2 ,e3'k. The ferromagnetic state is therefore stable atT.0 as well as at T,0, but the domain structures differ be-cause of the different relative orientations of ^I(k)& and k.Dipolar ferromagnetic ordering in an fcc lattice has recentlybeen investigated experimentally by Roser and Corruccini(1990) in several Cs 2NaR(NO 2) 6 rare-earth salts, and theo-retically by Bouchaud and Zerah (1993).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

tions in silver at T,0 have been illustrated in Fig. 127.There are xN and (12x)N spins in domains A and B,respectively. It is useful to express IA and IB in terms ofa static and an oscillating part,

IA ,B5Iav1IA ,B8 , (171)

where

Iav5^I~k50 !&5xIA1~12x !IB (172)

is the magnetization, apart from a constant coefficient.The oscillating parts are

IA8 5 (kÞ0

^I~k!&exp~ ik•riPA!5~12x !~IA2IB!, (173)

IB8 5 (kÞ0

^I~k!&exp~ ik•riPB!52x~IA2IB!. (174)

The local fields acting on spins in the two domains are

BiPA5B01@lz~0 !Iav1l1~k→0 !IA8 #/~\g!, (175)

BiPB5B01@lz~0 !Iav1l1~k→0 !IB8 #/~\g!, (176)

where the static external field B05B0z. Here we havemade use of Eq. (75) and the assumption that the widthof the domains is large in comparison with the latticeparameter, so that only small k-vectors are relevant (seep. 490 in Abragam and Goldman, 1982). There should,however, be many domains in the sample. We have fur-ther assumed that the oscillating component of the do-main magnetization, proportional to IA2IB , is alignedalong the direction of modulation, i.e., the nonzero wavevectors k in Eqs. (173) and (174). This selects the coef-ficients l1(k→0) which multiply IA ,B8 in the expressionsfor the local fields.

For a constant entropy and magnetic field, the stablestate corresponds to a maximum of the magneticenthalpy E . The entropy is constant if uIAu5uIBu=constant. The stable states can be found by maximizingthe enthalpy with respect to the division parameter xand the directions of IA and IB . Viertio and Oja (1992)derived an equation for E by using Eq. (78) and theexpressions for the local fields BiPA and BiPB . It ismore convenient, however, to make use of Eq. (84). As-suming that Iav is along the external field, and neglectingall contributions from domain walls, we obtain

E/N52\gB0Iav212 @l~0 !2l1~k→0 !#Iav

2

2 12 l1~k→0 !I2p2. (177)

FIG. 127. Three degenerate domain configurations of silvernuclei at T,0 in a magnetic field B pointing upwards. Modi-fied from Viertio and Oja (1992).

Page 126: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

126 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

We have here made use of the sum rule

Iav2 1 (

kÞ0u^I~k!&u25I2p2, (178)

where p is the polarization and uIAu5uIBu5pI . Themaximum of E is found by requiring ]E/]x5(]E/]Iav)(]Iav /]x)50. The nontrivial solution yieldsthe magnetization

M5\grIav52B0 /@m0~12Dz!# . (179)

It should be noted that M is antiparallel to the externalfield in order to maximize E . The domain structure isstable below

Bc5m0\grIp~12Dz!. (180)

In this sense there is a ferromagnetic transition in a non-vanishing external field.

The above equations yield local fields Bi5l1(k→0)^Ii&/(\g); here u^Ii&u5pI where p is ob-tained from Eq. (76). Thus the magnitude of the localfields and spins is indeed constant, as was assumed at thebeginning of the calculation. Hence the solutions ob-tained by maximizing E correspond to the stable state.

For the domain configuration illustrated in Fig.127(a), the magnetization depends on the relative vol-umes of the two types of domains. When the externalfield is increased, the domains with magnetizations anti-parallel to the field grow at the expense of those withparallel magnetizations. The equilibrium M of Eq. (179)corresponds to the relative volume of domains A ex-pressed by

x5 12 ~11B0 /Bc!. (181)

This kind of domain structure has been found in dielec-tric nuclear magnets both theoretically (Abragam andGoldman, 1982) and experimentally. Neutron-diffraction measurements by Roinel et al.. (1980) re-vealed the presence of ferromagnetic domains in theform of thin slices, approximately 20-Å thick, perpen-dicular to the external field.

Figures 127(b) and 127(c) show configurations that donot exist in dielectrics in the rotating frame. In the struc-ture of Fig. 127(b), spins adjust to the external field sim-ply by canting, and no movement of domain walls isnecessary. In this case, the relative volumes of the twodomains are always equal, i.e., x5 1

2. The structure inFig. 127(c) is a superposition of the patterns (a) and (b).Spins can adjust to the external field both by canting andby domain-wall movement. It is also possible to have adomain-modulation in the third Cartesian direction.

According to Viertio and Oja (1992), the generalcharacteristics of domain structures at T,0 are that thetangential component of magnetization is always con-tinuous and the normal component changes sign at adomain boundary. The opposite is true at positive spintemperatures.

2. Demagnetization into the domain state

The enthalpy per spin for the domain configurationsof Fig. 127 is given by

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

E/N5 12 B2/@m0r~12Dz!#2 1

2 l1~k→0 !I2p2. (182)

Figure 128 shows how E depends on the external fieldfor samples with different shapes. It is obvious that for-mation of domains makes it possible for, say, a sphere-shaped specimen to mimic the most advantageoussample shape at T,0, namely, an infinite plate perpen-dicular to the external field.

Since the local fields have a constant value for thedomain states, the structures are permanent (see Sec.XV.B.4). It then follows, similarly to the type-I fcc anti-ferromagnets (see Sec. XV.B.5 and Fig. 7), that the linesof constant entropy are vertical in the (T ,B) region cor-responding to the domain state.

Although the three domain configurations of Fig. 127are all degenerate in E , the processes that are needed tocreate them during demagnetization are very different.As illustrated in Fig. 129, the sample can be demagne-tized into the domain state of Fig. 127(b) by continuouscanting of the spins. The situation is completely differentfor the two other domain configurations illustrated inFig. 127. In case (a), when the system is demagnetizedbelow Bc , domains with magnetizations completely op-posite to the rest of the sample should be created tomaximize E . Nucleation of such domains is likely to be aslow process. Adjustment to the external field for do-main configuration (a) is also much more difficult than

FIG. 128. Magnetic enthalpy for samples of different shapes(infinite disc normal to B, sphere, and an infinitely long rodwith its axis along B) in an external magnetic field at T520.Domains are created below the respective critical fieldsBc

disc50, Bcrod5m0\grI , and Bc

sph523Bc

rod .

FIG. 129. Spin directions during demagnetization into a do-main state. For explanations, see text. Modified from Viertioand Oja (1992).

Page 127: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

127A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

for state (b) because movement of domain walls is cer-tainly a slower process than canting of individual spins.Similarly, we expect that configuration (c) suffers from aslow nucleation, although the problem is not so severeas for the state (a). We conclude that domain configura-tion (b) is the most likely to be present in experiments.

If the spins are indeed ordered in the domain patternof Fig. 127(b), it should be possible to reverse B with nohysteresis (Viertio and Oja, 1992). This is in agreementwith measurements on silver that did not show any hys-teresis while the field direction across the ferromagneti-cally ordered region was slowly reversed (Hakonen,Nummila, Vuorinen, and Lounasmaa, 1992).

3. Comparison with experimental data on silver

The ferromagnetic nuclear-spin structure in silver atT,0 has been investigated using Monte Carlo simula-tions (Viertio and Oja, 1992). In these calculations thedemagnetization factors were chosen as Dx5Dy50 andDz51 to imitate the equilibrium inside a single domain.The spin-spin interactions were summed up to theeighth neighboring cell. Simulations found ferromag-netic ordering along the z axis above T5TC521.7 nK.Adiabatic demagnetization was also simulated, wherebya critical initial polarization pc530% for ferromagneticordering was obtained. This is clearly lower than theobserved pc5(4965)%, although the theoretical TC isin good agreement with the measured TC52(1.960.4)nK.

These Monte Carlo simulations were made for classi-cal spins with magnetic moments m5g\I . Since nuclearspins in silver have I5 1

2, the assumption of a classicalspin is a severe approximation. As we discussed in Sec.XV.C.3c, a rule of thumb for taking into account thisdifference is to multiply the simulated TC by the factor(I11)/I . This leads, however, to almost a factor-of-3discrepancy between the theoretical and observed TC’sof silver.

Unlike the ordering temperature, static magnetic sus-ceptibility is expected to be rather insensitive to I . Thesimulations and the mean-field theory both reproducedthe measured result for the transversal susceptibility; thecorresponding three values fall between x520.9 andx521.1.

XVI. SUMMARY AND FUTURE PROSPECTS

A. Copper

The magnetic phase diagram of copper is summarizedin Fig. 130 on the basis of experimental and theoreticalstudies. The (0 2

323) and (1 0 0) spin configurations have

been investigated in detail by neutron-diffraction mea-surements and by theoretical calculations. The generaltrend is that the (1 0 0) order is found in low and highfields while the (0 2

323) spin configuration is stable in in-

termediate fields. An important exception however, is,the high-field configuration for the Bi[111] alignment,where no (1 0 0) order has been found. This remains an

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

open question although several suggestions have beenmade, including, in particular, an ordering vector of theform Q=(h ,k ,l), with uhu, uku, and ulu all unequal andnonzero. Another possibility is an incommensuration ofthe k=(p/a)(0, 2

3, 23) ordering vector along the @0hh# di-

rection.In the high-field regions for the [011] and [100] field

alignments, neutron-diffraction data were consistentwith the theoretically predicted triple-k and double-ktype-I structures. In low and intermediate fields, align-ment of B with respect to the crystalline axes had a dra-matic effect on the stability of various cubic-symmetry-related (0 2

323) Bragg reflections. The fact that the

measurements completely agreed with theoretical selec-tion rules adds confidence to the calculated spin configu-rations. In low intermediate fields, measurements foundsimultaneous (0 2

323) and (1 0 0) orders, which seemed to

be the result of superposition of the two ordering vec-tors in a single magnetic domain. The proposal must stillbe confirmed experimentally. For field alignments some-what off the [100] direction it is possible that a superpo-sition structure extends to relatively high fields, assketched in Fig. 130.

B. Other simple metals

The magnetic phase diagram of silver was summarizedin Fig. 4. The observation of antiferromagnetism at posi-tive temperatures and ferromagnetism at negative tem-peratures provides an excellent illustration of basic prin-ciples of statistical physics. At T.0, the type-I orderingvector has been found in recent neutron-diffractionmeasurements (Tuoriniemi, Nummila et al., 1995). Newinformation on ordered spin structures for differentalignments of the field will soon be available. The dis-

FIG. 130. Magnetic phase diagram of copper nuclei in theplane By5Bz . The illustration is based on a combination ofneutron-diffraction data and theoretical calculations. The grayarea corresponds to fields in which the (0 2

323) and (1 0 0) or-

ders exist simultaneously. The form of this region is tentative.The ordering vector Q=(h ,k ,l) is only one of several possibili-ties.

Page 128: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

128 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

crepancy between theoretical and observed orderingtemperatures, both at T.0 and T,0, warrants furtherinvestigations.

The obvious next goal in studies on rhodium is to ob-serve nuclear-spin ordering.

The origin of the high ferromagnetic ordering tem-perature of 115In in AuIn 2 is a subject of considerableinterest. To further substantiate the presence of a largeexchange interaction, it might be useful to study ex-change merging of the NMR lines of 115In and 113In ashas been done in thallium. Electronic band-structurecalculations of exchange forces are very much in order.The importance of magnetoelastic coupling betweennuclear spins and the lattice should be further investi-gated as has been discussed by Siemensmeyer andSteiner (1992).

There are several other simple metals for which onecan expect important progress in the study of nuclearordering. These include, for example, gold, thallium,scandium, platinum, and yttrium. The presence of mag-netic impurities, even at the level of one-ppm, is a prob-lem in several of these materials and poses seriousmaterials-science challenges. To estimate the orderingtemperature in these metals one may assume that Tc isdetermined by the Ruderman-Kittel mechanism. Ac-cording to the free-electron picture, the strength of theRuderman-Kittel interaction is approximately propor-tional to the inverse of the Korringa constant k (see,for example, Oja and Kumar, 1987; Slichter, 1990;Herrmannsdorfer and Pobell, 1995). One expects thenthat, within an order of magnitude, Tc'I(I11)/k ,where I is the spin. The observed Tc’s of copper, silver,and AuIn 2 display such a dependence, as shown in Fig.131. A rough estimate of the ordering temperature inAu, Rh, Tl, Sc, Pt, and Y can be obtained on the basis of

FIG. 131. Observed nuclear-ordering temperatures of silver,copper, and AuIn 2 vs I(I11)/k ; I is the spin, and k is theKorringa constant. The line Tc5531028 K [I(I+1)/k] sK isonly a guide for the eye. The solid circle represents the abso-lute value of the Curie temperature of silver at T,0. Thevalues for I(I11)/k in several other nuclear-spin systems havebeen indicated as well.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

their I(I11)/k parameters, which are indicated in thefigure as well.

C. Temperature records

Usually as a by-product of their research, ultralow-temperature physicists are, from time to time, drawingcloser to their elusive goal of achieving the absolutezero. Figure 132 (see p. 58) illustrates the progress dur-ing the last 25 years. What has been achieved dependson the substance that has been refrigerated.

3He is an important subject of research at ultralowtemperatures. This includes the superfluid liquid, dilutemixtures of 3He and 4He, and solid 3He. Superfluid3He has been refrigerated below 100 mK by the groupof George Pickett at Lancaster University (Carney et al.,1989), and more recently by the group of Frank Pobellat the University of Bayreuth (Konig, Betat, and Pobell,1994) and by Bunkov and Fisher (1995) at Grenoble.Dilute solutions have been cooled to 100 mK by theOsaka team (Oh et al., 1994) and by the group atBayreuth (Konig, Betat, and Pobell, 1994). Solid 3Hehas been refrigerated to temperatures around 30–40mK by groups at Tokyo, Osaka, and Nagoya (Takanoet al., 1985; Yano et al., 1990; Suzuki, Kondo et al., 1991;Okamoto et al., 1994).

Metals offer another important category of specimensthat have been studied at ultralow temperatures.Conduction-electron temperatures in copper of about 10mK and below have been measured by the groups ofPobell (Gloos et al., 1988) and Pickett (Enrico et al.,1994). In Helsinki, nuclear spins have been cooled to 280pK in rhodium, while the ‘‘hot’’ record at negativenuclear temperatures is 2750 pK (see Sec. X.A). Spon-taneous spin ordering has been observed at 560 pK forantiferromagnetism and at 21.9 nK for ferromagnetismin silver (see Sec. VIII).

The newest entry into the race for low-temperaturerecords is Bose-Einstein condensation in gaseous assem-blies of several alkaline atoms. Researchers at the Na-tional Institute of Standards and Technology in Boulder,together with their colleagues at the University of Colo-rado, first reported Bose-Einstein condensation in ru-bidium atoms near a temperature of 170 nanokelvin(Anderson et al., 1995). The progress in this field hasbeen very rapid since this breakthrough (Culotta, 1995).

ACKNOWLEDGMENTS

Preparation of this review was financially supportedby the Academy of Finland. We express our gratitude toour colleagues Pertti Hakonen, Marko Heinila, KimLefmann, Kaj Nummila, Hanna Viertio-Oja, Reko Vuo-rinen, and Juha Tuoriniemi for several useful discus-sions. We have also benefitted from comments by YuriyBunkov, Georg Eska, George Pickett, Frank Pobell,Martti Salomaa, Haruhiko Suzuki, Michael Steiner, andYasumasa Takano. We are indebted to Harri Colerus,Ursula Holmstrom, Tauno Knuuttila, and Juha Marti-kainen for help in preparing the illustrations. Finally, we

Page 129: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

129A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

thank the anonymous referees for their useful and de-tailed comments on the manuscript.

REFERENCES

Aalto, M. I., P. M. Berglund, H. K. Collan, G. J. Ehnholm, R.G. Gylling, and O. V. Lounasmaa, 1972, ‘‘Nuclear spin-latticerelaxation of the Zeeman and the spin-spin systems in copperbetween 1 and 17 mK,’’ Physica 61, 314–320.

Abragam, A., 1961, The Principles of Nuclear Magnetism(Clarendon, Oxford).

Abragam, A., 1987, ‘‘Ordering nuclear spins,’’ Proc. R. Soc.London, Ser. A 412, 255–268.

Abragam, A., and M. Goldman, 1982, Nuclear Magnetism: Or-der and Disorder (Clarendon, Oxford).

Abragam, A., J. F. Jacquinot, M. Chapellier, and M. Goldman,1973, ‘‘Absortion lineshape of highly polarized nuclear spinsystems,’’ J. Magn. Reson. 10, 322–346.

Abragam, A., and W. G. Proctor, 1958, ‘‘Spin temperature,’’Phys. Rev. 109, 1441–1458.

Akai, K., and H. Ishii, 1994, ‘‘Electronic specific heat in metal-lic Van Vleck paramagnets,’’ Phys. Rev. B 49, 275–284.

Akhiezer, A. I., V. G. Bar’yakhtar, and S. V. Peletminskii,1968, Spin Waves (North-Holland, Amsterdam).

Alloul, H., and C. Froidevaux, 1967, ‘‘Nuclear-magnetic-resonance spin echoes in alloys,’’ Phys. Rev. 163, 324–335.

Al’tshuler, S. A., 1966, ‘‘Use of substances containing rare-earth ions with even number of electrons to obtain infralowtemperatures,’’ Pis’ma Zh. Eksp. Teor. Fiz. 3, 180–183 [Sov.Phys. JETP Lett. 3, 112–115 (1966)].

Anderson, A. G., 1962, ‘‘Nuclear spin absortion spectra in sol-ids,’’ Phys. Rev. 125, 1517–1527.

Anderson, M. H., J. R. Ensher, M. R. Matthews, C. E. Wie-man, and E. A. Cornell, 1995, ‘‘Observation of Bose-Einsteincondensation in a dilute atomic vapor,’’ Science 269, 198–201.

Anderson, P. W., 1973, ‘‘Resonating valence bonds: A newkind of insulator?’’ Mater. Res. Bull. 8, 153–160.

Andres, K., and E. Bucher, 1968, ‘‘Observation of hyperfine-enhanced nuclear magnetic cooling,’’ Phys. Rev. Lett. 21,1221–1223.

Andres, K., and E. Bucher, 1972, ‘‘Nuclear cooling inPrCu6,’’ J. Low Temp. Phys. 9, 267–289.

Andres, K., E. Bucher, J. P. Maita, and A. S. Cooper, 1972,‘‘Observation of cooperative nuclear magnetic order inPrCu2 below 54 mK,’’ Phys. Rev. Lett. 28, 1652–1655.

Andres, K., E. Bucher, P. H. Schmidt, J. P. Maita, and S.Darack, 1975, ‘‘Nuclear-induced ferromagnetism below 50mK in the Van Vleck paramagnet PrCu 5,’’ Phys. Rev. B 11,4364–4372.

Andres, K., E. Hagn, E. Smolic, and G. Eska, 1975, ‘‘APrCu6 cryostat for nuclear orientation experiments down to2.7 mK,’’ J. Appl. Phys. 46, 2752–2759.

Andres, K., and O. V. Lounasmaa, 1982, ‘‘Recent progress innuclear cooling,’’ Prog. Low Temp. Phys. 8, 221–287.

Andrew, E. R., 1973, ‘‘NMR in rapidly-rotated metals,’’ inProceedings of the XVII Congress Ampere, edited by V. Hovi(North-Holland, Amsterdam), pp. 18–32.

Andrew, E. R., J. L. Carolan, and P. J. Randall, 1971, ‘‘Mea-surement of the Ruderman-Kittel interaction for copper,’’Phys. Lett. A 37, 125–126.

Andrew, E. R., and W. S. Hinshaw, 1973, ‘‘Indirect nuclearinteraction coupling constants for metallic copper,’’ Phys.Lett. A 43, 113–114.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Angerer, U., and G. Eska, 1984, ‘‘Nuclear demagnetization ofthallium,’’ Cryogenics 53, 515–521.

Annila, A. J., K. N. Clausen, P.-A. Lindgard, O. V. Lounas-maa, A. S. Oja, K. Siemensmeyer, M. Steiner, J. T. Tuorini-emi, and H. Weinfurter, 1990, ‘‘Nuclear order in copper: Newtype of antiferromagnetism in an ideal fcc system,’’ Phys.Rev. Lett. 64, 1421–1424.

Annila, A. J., K. N. Clausen, A. S. Oja, J. T. Tuoriniemi, andH. Weinfurter, 1992, ‘‘Neutron-diffraction studies of thenuclear magnetic phase diagram of copper,’’ Phys. Rev. B 45,7772–7788.

Asada, T., K. Terakura, and T. Jarlborg, 1981, ‘‘An analysis ofthe spin-lattice relaxation of cubic transition metals,’’ J. Phys.F 11, 1847–1857.

Ashcroft, N. W., and N. D. Mermin, 1976, Solid State Physics(Holt, Rinehart, and Winston, New York).

Babcock J., J. Kiely, T. Manley, and W. Weyhmann, 1979,‘‘Nuclear magnetic ordering in PrCu 6,’’ Phys. Rev. Lett. 43,380–383.

Bacon, F., J. A. Barclay, W. D. Brewer, D. A. Shirley, J. E.Templeton, 1972, ‘‘Temperature-independent spin-lattice re-laxation time in metals at very low temperatures,’’ Phys. Rev.B 5, 2397–2409.

Ballou, R., P. J. Brown, J. Deportes, A. S. Markosyan, and B.Ouladdiaf, 1992, ‘‘Exchange frustration and metastability ofthe magnetic structure of TbMn 2,’’ J. Magn. Magn. Mater.104–107, 935–936.

Ballou, R., J. Deportes, R. Lemaire, and B. Ouladdiaf, 1988,‘‘Mn magnetism and magnetic structures in RMn 2,’’ J. Appl.Phys. 63, 3487–3489.

Bardeen, J., 1937, ‘‘An improved calculation of the energies ofmetallic Li and Na,’’ J. Chem. Phys. 6, 367–371.

Bauml, W., G. Eska, and W. Pesch, 1994, ‘‘NMR in metals atlow temperatures with particular emphasis on the finite skindepth,’’ Physica B 194–196, 321–322.

Benoit, A., J. Bossy, J. Flouquet, and J. Schweizer, 1985,‘‘Magnetic diffraction in solid 3He,’’ J. Phys. (Paris) Lett. 46,923–927.

Benoit, A., J. Flouquet, J. L. Genicon, and J. Palleau, 1981,‘‘Magnetic ordering in PrCu 5 by neutron diffaction,’’ PhysicaB 108, 1103–1104.

Benoit, A., J. Flouquet, D. Rufin, and J. Schweizer, 1982a,‘‘Magnetic transition of solid 3He observed by polarized neu-trons,’’ J. Phys. (Paris) Lett. 43, 431–436.

Benoit, A., J. Flouquet, D. Rufin, and J. Schweizer, 1982b,‘‘Thermometry at very low temperatures with polarized neu-trons; application to the search for the 3He magnetic struc-ture,’’ J. Phys. (Paris) Colloq. 43, C7, 311–316.

Berglund, P. M., H. K. Collan, G. J. Ehnholm, R. G. Gylling,and O. V. Lounasmaa, 1972, ‘‘The design and use of nuclearorientation thermometers employing 54Mn and 60Co nuclei inferromagnetic hosts,’’ J. Low Temp. Phys. 6, 357–383.

Berlin, T. H., and M. Kac, 1952, ‘‘The spherical model of aferromagnet,’’ Phys. Rev. 86, 821–835.

Betts, D. S., 1976, Refrigeration and Thermometry below OneKelvin (Sussex University Press, London).

Binder, K., J. L. Lebowitz, M. K. Phani, and M. H. Kalos, 1981,‘‘Monte Carlo study of the phase diagrams of binary alloyswith face centered cubic lattice structure,’’ Acta Metall. 29,1655–1665.

Binder, K., and A. P. Young, 1986, ‘‘Spin glasses: Experimen-tal facts, theoretical concepts, and open questions,’’ Rev.Mod. Phys. 58, 801–976.

Page 130: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

130 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Bloembergen, N., and T. J. Rowland, 1955, ‘‘Nuclear spin ex-change in solids: Tl 203 and Tl 205 magnetic resonance in thal-lium and thallic oxide,’’ Phys. Rev. 97, 1679–1698.

Bloembergen, N., S. Shapiro, P. S. Pershan, and J. O. Artman,1959, ‘‘Cross-relaxation in spin systems,’’ Phys. Rev. 114,445–459.

Bloembergen, N., and S. Wang, 1954, ‘‘Relaxation effects inpara-and ferromagnetic resonance,’’ Phys. Rev. 93, 72–83.

Bouchaud, J. P., and P. G. Zerah, 1993, ‘‘Dipolar ferromag-netism: A Monte Carlo study,’’ Phys. Rev. B 47, 9095–9097.

Bouffard, V., C. Fermon, J. F. Gregg, J. F. Jacquinot, and Y.Roinel, 1994, ‘‘Nuclear magnetic ordering ‘avant toutechose’,’’ in NMR and More in Honour of Anatole Abragam,edited by M. Goldman and M. Porneuf (Les Editions de Phy-sique, Les Ulis), pp. 81–102.

Bradley, D. I., A. M. Guenault, V. Keith, C. J. Kennedy, I. E.Miller, S. G. Mussett, G. R. Pickett, and W. P. Pratt Jr., 1984,‘‘New methods for nuclear cooling into the microkelvin re-gime,’’ J. Low Temp. Phys. 57, 359–390.

Bucal, Ch., K. J. Fischer, M. Kubota, R. M. Mueller, and F.Pobell, 1978, ‘‘Nuclear demagnetization of PrS and PrNi5,’’ J.Phys. (Paris) 39, L-457–L-458.

Buchal, C., J. Hanssen, R. M. Mueller, and F. Pobell, 1978,‘‘Platinum wire NMR thermometer for ultralow tempera-tures,’’ Rev. Sci. Instrum. 49, 1360–1361.

Bucal, Ch., R. M. Mueller, F. Pobell, M. Kubota, and H. R.Folle, 1982, ‘‘Superconductivity investigations of Au-In alloysat ultralow temperatures,’’ Solid State Commun. 42, 43–47.

Buchal, C., F. Pobell, R. M. Mueller, M. Kubota, and J. R.Owers-Bradley, 1983, ‘‘Superconductivity of rhodium at ul-tralow temperatures,’’ Phys. Rev. Lett. 50, 64–67.

Buishvili, L. L., and N. P. Fokina, 1994, ‘‘Cross relaxation inweak magnetic fields at ultralow spin temperatures,’’ Zh.Eksp. Teor. Fiz. 106, 627–632 [Sov. Phys. JETP 79, 344–346(1994)].

Buishvili, L. L., D. A. Kostarov, and N. P. Fokina, 1994,‘‘Merging-splitting and suppression-enhancement in magneticresonance lines of highly polarized exchange-coupled non-equivalent spins,’’ Zh. Eksp. Teor. Fiz. 106, 1477–1488 [J.Expt. Theor. Phys. 79, 799–894 (1994)].

Bunkov, Yu. M., 1989, ‘‘Superconducting aluminium heatswitch prepared by diffusion welding,’’ Cryogenics 29, 938–939.

Bunkov, Yu. M., and S. Fisher, 1995, ‘‘NMR in superfluid3He at very low temperatures,’’ J. Low Temp. Phys. 101, 123–134.

Carney J. P., A. M. Guenault, G. R. Pickett, and G. F. Spencer,1989, ‘‘Extreme nonlinear damping of the quasiparticle gas insuperfluid 3He-B in the low-temperature limit,’’ Phys. Rev.Lett. 62, 3042–3045.

Casimir, H. B. G., and F. K. Du Pre, 1938, ‘‘Note on the ther-modynamic interpretation of paramagnetic relaxation phe-nomena,’’ Physica 5, 507–511.

Chamberlin, R. V., L. A. Moberly, and O. G. Symko, 1979,‘‘High-sensitivity magnetic resonance by SQUID detection,’’J. Low Temp. Phys. 35, 337–347.

Chapellier, M., M. Goldman, V. H. Chau, and A. Abragam,1969, ‘‘Production et observation d’un etat antiferromagnet-ique nucleaire,’’ C.R. Acad. Sci. 268, 1530–1533.

Chapellier, M., M. Goldman, V. H. Chau, and A. Abragam,1970, ‘‘Production and observation of a nuclear antiferromag-netic state,’’ J. Appl. Phys. 41, 849–853.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Chapman, A. C., P. Rhodes, and E. F. Seymour, 1957, ‘‘Theeffect of eddy currents on nuclear magnetic resonance in met-als,’’ Proc. Phys. Soc. London 70, 345–360.

Chen, K., and D. P. Landau, 1994, ‘‘Spin-dynamics study of thedynamic critical behavior of the three-dimensional classicalHeisenberg ferromagnet,’’ Phys. Rev. B 49, 3266–3274.

Clausen, K. N., 1991, ‘‘Neutron scattering at nK tempera-tures,’’ Neutron News 2, 25–28.

Cohen, M. H., and F. Keffer, 1955, ‘‘Dipolar sums in the primi-tive cubic lattices,’’ Phys. Rev. 99, 1128–1134.

Cross, M. C., and D. S. Fisher, 1985, ‘‘Magnetism in solid3He: confrontation between theory and experiment,’’ Rev.Mod. Phys. 57, 881–921.

Culotta, E., 1995, ‘‘A new form of matter unveiled,’’ editorialarticle in Science 270, 1902–1903.

De Jongh, L. J., and A. R. Miedema, 1974, ‘‘Experiments onsimple magnetic model systems,’’ Adv. Phys. 23, 1–260.

De Klerk, D., 1956, ‘‘Adiabatic demagnetization,’’ in Encyclo-pedia of Physics (Springer, Berlin), Vol. 15, pp. 38–209.

Deimling, M., H. Brunner, J. P. Colpa, K. P. Dinse, and K. H.Hausser, 1980, ‘‘Microwave-induced optical nuclear polariza-tion (MI-ONP),’’ J. Magn. Reson. 39, 185–202.

Demco, D. E., J. Tegenfeldt, and J. S. Waugh, 1975, ‘‘Dynam-ics of cross relaxation in nuclear magnetic double resonance,’’Phys. Rev. B 11, 4133-4151.

Diep, H. T., and H. Kawamura, 1989, ‘‘First-order phase tran-sition in the fcc Heisenberg antiferromagnet,’’ Phys. Rev. B40, 7019–7022.

Domb, C., and A. R. Miedema, 1964, ‘‘Magnetic transitions,’’in Progress in Low Temperature Physics, edited by C. J.Gorter (North-Holland, Amsterdam), Vol. 4, pp. 296–343.

Ebert, H., H. Winter, and J. Voitlander, 1984, ‘‘Nuclear spinlattice relaxation rates of Ag in AgPd and Cu in CuPd,’’ J.Phys. F 14, 2433–2442.

Ehnholm, G. J., J. P. Ekstrom, J. F. Jacquinot, M. T. Loponen,O. V. Lounasmaa, and J. K. Soini, 1979, ‘‘Evidence fornuclear antiferromagnetism in copper,’’ Phys. Rev. Lett. 42,1702–1705.

Ehnholm, G. J., J. P. Ekstrom, J. F. Jacquinot, M. T. Loponen,O. V. Lounasmaa, and J. K. Soini, 1980, ‘‘NMR studies onnuclear ordering in metallic copper below 1 5 K,’’ J. LowTemp. Phys. 39, 417–450.

Ehnholm, G. J., J. P. Ekstrom, M. T. Loponen, and J. K. Soini,1979, ‘‘Transversal SQUID NMR,’’ Cryogenics 19, 673–678.

Ehrlich, A. C., 1974, ‘‘Oxygen annealing of silver for obtaininglow electrical resistivity: technique and interpretation,’’ J.Mater. Sci. 9, 1064–1072.

Ekstrom, J. P., J. F. Jacquinot, M. T. Loponen, J. K. Soini, andP. Kumar, 1979, ‘‘Nuclear spin interaction in copper: NMR athigh polarization and in low fields,’’ Physica B 98, 45–52.

Enrico, M. P., S. N. Fisher, A. M. Guenault, I. E. Miller, andG. R. Pickett, 1994, ‘‘Temperature dependence of the nuclearspin-lattice relaxation time in copper metal to below 10mK,’’ Phys. Rev. B 49, 6339–6342.

Eska, G., 1989, ‘‘Nonlinear spin dynamics and nuclear order-ing: Tl metal as an example,’’ in Symposium on QuantumFluids and Solids—1989, AIP Conference Proceedings No.194, edited by G. G. Ihas and Y. Takano (AIP, New York),pp. 316–327.

Eska, G., and E. Schuberth, 1987, ‘‘NMR behavior of Tl metalat very low temperatures,’’ Jpn. J. Appl. Phys. Suppl. 26-3,435–436.

Page 131: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

131A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Eska, G., E. Schuberth, and B. Turrell, 1986, ‘‘Nuclear spin-lattice relaxation of thallium at low temperatures,’’ Phys.Lett. A 115, 413–416.

Fickett, F. R., 1974, ‘‘Oxygen annealing of copper: A review,’’Mater. Sci. Eng. 14, 199–210.

Foner, S. 1963, ‘‘Antiferromagnetic and ferrimagnetic reso-nance,’’ in Magnetism, edited by G. T. Rado and H. Suhl(Academic, New York), Vol. I, pp. 383–447.

Frisken, S. J., 1989, Ph.D. Thesis (University of New SouthWales).

Frisken, S. J., and D. J. Miller, 1986, ‘‘Realistic calculation ofthe indirect-exchange interaction in metals,’’ Phys. Rev. Lett.57, 2971–2974.

Frisken, S. J., and D. J. Miller, 1988a, ‘‘Monte Carlo study ofantiferromagnetic nuclear ordering in Cu,’’ Phys. Rev. Lett.61, 1017–1020.

Frisken, S. J., and D. J. Miller, 1988b, ‘‘Indirect-exchange in-teraction in copper,’’ Phys. Rev. B 37, 10884–10886.

Froidevaux, C., and M. Weger, 1964, ‘‘Direct measurement ofthe Ruderman-Kittel interaction in platinum alloys,’’ Phys.Rev. Lett. 12, 123–125.

Garrett, C. G. B., 1951, ‘‘Experiments with an anisotropicmagnetic crystal at temperatures below 1K,’’ Proc. R. Soc.London, Ser. A 206, 242–257.

Gehring, G. A. and K. Z. Gehring, 1975, ‘‘Co-operative Jahn-Teller effects,’’ Rep. Prog. Phys. 38, 1–89.

Genicon, J. L., J. L. Tholence, and R. Tournier, 1978, ‘‘Nuclearinduced ferromagnetism in PrCu 5,’’ J. Phys. (Paris) Colloq.39, C6, 798–799.

Gloos, K., R. Konig, P. Smeibidl, and F. Pobell, 1990,‘‘Anomalous nuclear magnetic properties of 115In inAuIn2,’’ Europhys. Lett. 12, 661–666.

Gloos, K., P. Smeibidl, C. Kennedy, A. Singsaas, P. Sekowski,R. M. Mueller, and F. Pobell, 1988, ‘‘The Bayreuth nucleardemagnetization refrigerator,’’ J. Low Temp. Phys. 73, 101–136.

Gloos, K., P. Smeibidl, and F. Pobell, 1991, ‘‘Nuclear spin–electron coupling of copper at microkelvin temperatures,’’ Z.Phys. B 82, 227–231.

Goldman, M., 1970, Spin Temperature and Nuclear MagneticResonance in Solids (Clarendon, Oxford).

Greywall, D. S., 1985, ‘‘ 3He melting-curve thermometry at mil-likelvin temperatures,’’ Phys. Rev. B 31, 2675–2683.

Griessen, R., M. Landolt, and H. R. Ott, 1971, ‘‘A new anti-ferromagnetic phase in EuSe below 1.8 K,’’ Solid State Com-mun. 9, 2219–2223.

Griffiths, J. H. E., J. Owen, J. G. Park, and M. F. Partridge,1959, ‘‘Exhange interactions in antiferromagnetic salts of iri-dium I. Paramagnetic resonance experiments,’’ Proc. R. Soc.London, Ser. A 250, 84–96.

Gukasov, A. G. Th. Bruckel, B. Dorner, V. P. Plakhty, W.Prandl, E. F. Shender, and O. P. Smirnov, 1988, ‘‘Quantumexchange magnon gap in an antiferromagnetic with dynami-cally interacting spin subsystems,’’ Europhys. Lett. 7, 83.

Gylling, R. G., 1971, ‘‘Construction and operation of a nuclearrefrigeration cryostat,’’ Acta Polytech. Scand. Ph. 81, xx–xx.

Hakonen, P. J., 1993, ‘‘Nuclear magnetic ordering in silver atpositive and negative spin temperatures,’’ Phys. Scr. T49,327–332.

Hakonen, P. J., 1994, ‘‘Calculation of nuclear-spin entropy insilver and rhodium at positive and negative temperatures us-ing Monte Carlo simulations,’’ Phys. Rev. B 49, 15363–15365.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Hakonen, P. J., and O. V. Lounasmaa, 1994, ‘‘Negative abso-lute temperatures: ‘Hot’ spins in spontaneous magnetic or-der,’’ Science 265, 1821–1825.

Hakonen, P. J., O. V. Lounasmaa, and A. S. Oja, 1991, ‘‘Spon-taneous nuclear magnetic ordering in copper and silver atnano and picokelvin temperatures,’’ J. Magn. Magn. Mater.100, 394–412.

Hakonen, P. J., K. K. Nummila, and R. T. Vuorinen, 1992,‘‘Spin dynamics in highly polarized silver at negative absolutetemperatures,’’ Phys. Rev. B 45, 2196–2200.

Hakonen, P. J., K. K. Nummila, R. T. Vuorinen, and O. V.Lounasmaa, 1992, ‘‘Observation of nuclear ferromagnetic or-dering in silver at negative nanokelvin temperatures,’’ Phys.Rev. Lett. 68, 365–368.

Hakonen, P. J., and R. T. Vuorinen, 1992, ‘‘Nuclear ferromag-netic ordering in silver at negative nanokelvin temperatures,’’J. Low Temp. Phys. 89, 177–186.

Hakonen, P. J., R. T. Vuorinen, and J. E. Martikainen, 1993,‘‘Nuclear antiferromagnetism in rhodium metal at positiveand negative nanokelvin temperatures,’’ Phys. Rev. Lett. 70,2818–2821.

Hakonen, P. J., R. T. Vuorinen, and J. E. Martikainen, 1994,‘‘Anomalous spin-lattice relaxation in dilute RhFe at positiveand negative nanokelvin temperatures,’’ Europhys. Lett. 25,551–556.

Hakonen, P. J., and S. Yin, 1991, ‘‘Investigations of nuclearmagnetism in silver down to picokelvin temperatures. II,’’ J.Low Temp. Phys. 85, 25–64.

Hakonen, P. J., S. Yin, and O. V. Lounasmaa, 1990, ‘‘Nuclearmagnetism in silver at positive and negative absolute tem-peratures in the low nanokelvin range,’’ Phys. Rev. Lett. 64,2707–2710.

Hakonen, P. J., S. Yin, and K. K. Nummila, 1991, ‘‘Phase dia-gram and NMR studies of antiferromagnetically orderedpolycrystalline silver,’’ Europhys. Lett. 15, 677–682.

Harmon, B. N., and X.-W. Wang, 1987, ‘‘Comment on ‘Real-istic calculation of the indirect-exchange interaction in met-als’,’’ Phys. Rev. Lett. 59, 379.

Harmon, B. N., X.-W. Wang, and P.-A. Lindgard, 1992, ‘‘Cal-culation of the Ruderman-Kittel interaction and the nuclearmagnetic ordering in silver,’’ J. Magn. Magn. Mater. 104–107,2113–2115.

Hastings, J. M., N. Elliott, and L. M. Corliss, 1959, ‘‘Antifer-romagnetic structures of MnS2, MnSe 2, and MnTe 2,’’ Phys.Rev. 115, 13–17.

Heinila, M. T., and A. S. Oja, 1993a, ‘‘Selection of the groundstate in type-I fcc antiferromagnets in an external magneticfield,’’ Phys. Rev. B 48, 7227–7237.

Heinila, M. T., and A. S. Oja, 1993b, ‘‘Long-range order pro-duced by the interaction between spin waves in classical fccHeisenberg models,’’ Phys. Rev. B 48, 16514–16523.

Heinila, M. T., and A. S. Oja, 1994a, ‘‘Breaking of degeneracyin classical Heisenberg antiferromagnets,’’ Physica B 194–196, 237–238.

Heinila, M. T., and A. S. Oja, 1994b, ‘‘Isotropic type I fccantiferromagnet in an external field,’’ Physica B 194–196,239–240.

Heinila, M. T., and A. S. Oja, 1994c, ‘‘Anisotropic magneticnearest-neighbor exchange interaction in the pyrite structure:MnTe 2, MnS 2, and MnSe 2,’’ Phys. Rev. B 49, 11995– 12002.

Heinila, M. T., and A. S. Oja, 1994d, ‘‘Simulation study ofNMR lineshapes in a classical nuclear spin system with twoisotopes,’’ Phys. Rev. B 50, 15843–15851.

Page 132: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

132 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Heinila, M. T., and A. S. Oja, 1995, ‘‘First-principles calcula-tion of antiferromagnetic NMR absorption curves,’’ Phys.Rev. B 52, R6967–R6970.

Heinila, M. T., and A. S. Oja, 1996, ‘‘Magnetic resonance oftype-I fcc antiferromagnets,’’ Phys. Rev. B 54, 9275–9287.

Henley, C. L., 1987, ‘‘Ordering by disorder: ground-state selec-tion in fcc vector antiferromagnets,’’ J. Appl. Phys. 61, 3962–3964.

Henley, C. L., 1989, ‘‘Ordering due to disorder in a frustratedvector antiferromagnet,’’ Phys. Rev. Lett. 62, 2056–2059.

Herrmannsdorfer, T., and F. Pobell, 1995, ‘‘Spontaneousnuclear ferromagnetic ordering of In nuclei in AuIn 2: Part I:Nuclear specific heat and nuclear susceptibility,’’ J. LowTemp. Phys. 100, 253–279.

Herrmannsdorfer, T., P. Smeibidl, B. Schroder-Smeibidl, andF. Pobell, 1995, ‘‘Spontaneous nuclear ferromagnetic order-ing of In in AuIn 2,’’ Phys. Rev. Lett. 74, 1665–1668.

Herrmannsdorfer, T., H. Uniewski, and F. Pobell, 1994a,‘‘Nuclear ferromagnetic ordering of 141Pr in the diluted VanVleck paramagnets Pr 12xY xNi 5,’’ Phys. Rev. Lett. 72, 148–151.

Herrmannsdorfer, T., H. Uniewski, and F. Pobell, 1994b,‘‘Nuclear ferromagnetic ordering of 141Pr in the diluted VanVleck paramagnets Pr 12xY xNi 5,’’ J. Low Temp. Phys. 97,189–211.

Hobden, M. V., and N. Kurti, 1959, ‘‘Experiments on nuclearcooling,’’ Philos. Mag. 4, 1092–1095.

Holstein, T., and H. Primakoff, 1940, ‘‘Field dependence of theintrinsic domain magnetization of a ferromagnet,’’ Phys. Rev.58, 1098–1113.

Huiku, M. T., 1984, ‘‘Nuclear magnetism in copper atnanokelvin temperatures and in low external magneticfields,’’ Physica B 126, 51–61.

Huiku, M. T., T. A. Jyrkkio, J. M. Kyynarainen, M. T.Loponen, O. V. Lounasmaa, and A. S. Oja, 1986, ‘‘Investiga-tions of nuclear antiferromagnetic ordering in copper atnanokelvin temperatures,’’ J. Low Temp. Phys. 62, 433–487.

Huiku, M. T., T. A. Jyrkkio, J. M. Kyynarainen, A. S. Oja, andO. V. Lounasmaa, 1984, ‘‘Phase diagram for spontaneousnuclear magnetic ordering in copper,’’ Phys. Rev. Lett. 53,1692–1695.

Huiku, M. T., T. A. Jyrkkio, and M. T. Loponen, 1983, ‘‘Ex-perimental curves of entropy and susceptilibity versus tem-perature for copper nuclear spins down to the ordered state,’’Phys. Rev. Lett. 50, 1516–1519.

Huiku, M. T., and M. T. Loponen, 1982, ‘‘Observation of amagnetic phase transition in the nuclear spin system of me-tallic copper at nanokelvin temperatures,’’ Phys. Rev. Lett.49, 1288–1291.

Huiku, M. T., M. T. Loponen, T. A. Jyrkkio, J. M. Kyy-narainen, A. S. Oja, and J. K. Soini, 1984, ‘‘Impurities and theanomalous spin-lattice relaxation in copper at submillikelvintemperatures,’’ in Proceedings of the 17th International Con-ference on Low Temperature Physics, edited by U. Eckern etal. (North-Holland, Amsterdam), pp. 133–134.

Huiku, M. T., and J. K. Soini, 1983, ‘‘Further studies of nuclearordering in metallic copper,’’ J. Low Temp. Phys. 50, 523–543.

Huiskamp, W. J., and O. V. Lounasmaa, 1973, ‘‘Ultralowtemperatures—how and why,’’ Rep. Prog. Phys. 36, 423–496.

Ishii, H., and S. Aoyama, 1991, ‘‘Quantum effect in enhancednuclear magnets,’’ Prog. Theor. Phys. Suppl. 106, 109–117.

Ishimoto, H., N. Nishida, T. Furubayashi, M. Shinohara, Y.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Takano, Y. Miura, and K. Ono, 1984, ‘‘Two-stage nucleardemagnetization refrigerator reaching 27 mK,’’ J. Low Temp.Phys. 55, 17– 31.

Jauho, P., and P. V. Pirila, 1970, ‘‘Spin-lattice relaxation ofnuclei due to electrons at very low temperatures,’’ Phys. Rev.B 1, 21–24.

Jeener, J., H. Eisendrath, and R. Van Steenwinkel, 1964,‘‘Thermodynamics of spin systems in solids,’’ Phys. Rev. 133,478A–490A.

Jensen, J., and A. R. Mackintosh, 1991, Rare Earth Magnetism(Clarendon, Oxford).

Jyrkkio, T. A., M. T. Huiku, K. N. Clausen, K. Siemensmeyer,K. Kakurai, and M. Steiner, 1988, ‘‘Calibration and applica-tions of polarized neutron thermometry at milli- and mi-crokelvin temperatures,’’ Z. Phys. B 71, 139–148.

Jyrkkio, T. A., M. T. Huiku, O. V. Lounasmaa, K. Siemens-meyer, K. Kakurai, M. Steiner, K. N. Clausen, and J. K.Kjems, 1988, ‘‘Observation of nuclear antiferromagnetic or-der in copper by neutron diffraction at nanokelvin tempera-tures,’’ Phys. Rev. Lett. 60, 2418–2421.

Jyrkkio, T. A., M. T. Huiku, K. Siemensmeyer, and K. N.Clausen, 1989, ‘‘Neutron diffraction studies of nuclear mag-netic ordering in copper,’’ J. Low Temp. Phys. 74, 435–473.

Karaki, Y., M. Kubota, and H. Ishimoto, 1994, ‘‘Specific heatof indium below 1 mK,’’ Physica B 194–196, 461–462.

Karimov, Yu. S., and I. F. Shchegolev, 1961, ‘‘Nuclear mag-netic resonance in metallic thallium,’’ Zh. Eksp. Teor. Fiz. 41,1082–1090 [Sov. Phys. JETP 14, 772–778 (1962)].

Kasevich, M., and S. Chu, 1992, ‘‘Laser cooling below a photonrecoil with three-level atoms,’’ Phys. Rev. Lett. 69, 1741–1744.

Kasuya, T., 1956, ‘‘A theory of metallic ferro-and antiferro-magnetism on Zener’s model,’’ Prog. Theor. Phys. 16, 45–57.

Kawarazaki, S., N. Kunitomi, J. R. Arthur, R. M. Moon, W. G.Stirling, and K. A. McEwen, 1988, ‘‘Evidence for nuclear-spinorder in double-hcp praseodymium by neutron diffraction,’’Phys. Rev. B 37, 5336–5341.

Keffer, F., 1966, ‘‘Spin waves,’’ in Handbuch der Physik(Springer, Berlin) Vol. 18/2, pp. 1–273.

Kittel, C., 1968, ‘‘Indirect exchange interactions in metals,’’ inSolid State Physics, Vol. 22 (Academic, New York), pp. 1–26.

Kittel, C., 1971, Introduction to Solid State Physics (Wiley,New York).

Kjaldman, L. H., P. Kumar, and M. T. Loponen, 1981,‘‘Nuclear magnetism in metals: Thermodynamics and mag-netic resonance,’’ Phys. Rev. B 23, 2051–2059.

Kjaldman, L. H., and J. Kurkijarvi, 1979, ‘‘Antiferromagneticordering of Cu-nuclei,’’ Phys. Lett. A 71, 454–456.

Knoll, G., 1968, Radiation Detection and Measurement (Wiley,New York).

Koester, L., and H. Rauch, 1981, Summary of Neutron Scatter-ing Lengths, IAEA Contract 2517/RB.

Koike, Y., H. Suzuki, S. Abe, Y. Karaki, M. Kubota, and H.Ishimoto, 1995, ‘‘Ferromagnetic nuclear spin order of scan-dium,’’ J. Low Temp. Phys. 101, 617–622.

Kolac, M., B. S. Neganov, and S. Sahling, 1985, ‘‘Low tempera-ture heat release from copper: Ortho-para conversion of hy-drogen,’’ J. Low Temp. Phys. 59, 547–559.

Komori, K., T. Goto, and S. Kobayashi, 1986, ‘‘Depolarizationof nuclear magnetization in aluminum small particles at lowtemperature,’’ J. Phys. Soc. Jpn. 55, 2133–2136.

Page 133: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

133A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Konig, R., A. Betat, and F. Pobell, 1994, ‘‘Refrigeration andthermometry of liquid 3He- 4He mixtures in the ballistic re-gime,’’ J. Low Temp. Phys. 97, 311–333.

Korringa, J., 1950, ‘‘Nuclear magnetic relaxation and reso-nance line shift in metals,’’ Physica 16, 601–610.

Kotzler, J., 1984, ‘‘On the possibility of fluctuation-driven first-order transitions,’’ Z. Phys. B 55, 119–129.

Kubo, R., and K. Tomita, 1954, ‘‘A general theory of magneticresonance absorption,’’ J. Phys. Soc. Jpn. 9, 888–919.

Kubota, M., K. J. Fischer, and R. M. Mueller, 1987, ‘‘Nonfer-romagnetic nuclear magnetic ordering in metallic hyperfineenhanced nuclear magnet PrSe,’’ Jpn. J. Appl. Phys. Suppl.26-3, 427–428.

Kubota, M., H. R. Folle, C. Buchal, R. M. Mueller, and F.Pobell, 1980, ‘‘Nuclear magnetic ordering in PrNi 5 at 0.4mK,’’ Phys. Rev. Lett. 45, 1812–1815.

Kubota, M., R. M. Mueller, C. Buchal, H. Chocholacs, J. R.Owers-Bradley, and F. Pobell, 1983, ‘‘Evidence for momentfluctuation effects in the metallic nuclear magnet PrNi 5,’’Phys. Rev. Lett. 51, 1382–1385.

Kubota, M., R. M. Mueller, and K. J. Fischer, 1984, ‘‘Study ofmagnetic ordering in metallic Pr compounds: Is simplest can-didate PrS?,’’ in Proceedings of the 17th International Confer-ence on Low Temperature Physics, edited by U. Eckern et al.(North-Holland, Amsterdam), pp. 15–16.

Kumar, P., J. Kurkijarvi, and A. S. Oja, 1985, ‘‘Phase diagramfor spontaneous nuclear magnetic ordering in copper,’’ Phys.Rev. B 31, 3194–3195.

Kumar, P., J. Kurkijarvi, and A. S. Oja, 1986, ‘‘Nuclear mag-netism in a metal,’’ Phys. Rev. B 33, 444–449.

Kumar, P., M. T. Loponen, and L. H. Kjaldman, 1980, ‘‘Spin-density-wave fluctuations in copper nuclear spins,’’ Phys.Rev. Lett. 44, 493–496.

Kurkijarvi, J., 1984, ‘‘Magnetic ordering of the nuclear spins inmetallic copper at nanokelvin temperatures and low appliedfields,’’ Physica B 127, 317–321.

Kurti, N., 1978, ‘‘From Cailletet and Pictet to microkelvin,’’Cryogenics 18, 451–458.

Kurti, N., 1982, ‘‘From the first mist of liquid oxygen to nuclearordering: anecdotes from the history of refrigeration,’’Physica B-C109-110, 1737–1752.

Kurti, N., F. N. Robinson, F. E. Simon, and D. A. Spohr, 1956,‘‘Nuclear cooling,’’ Nature 178, 450–453.

Landau, L. D., and E. M. Lifshitz, 1980, Statistical Physics, PartI (Pergamon, London).

Lefmann, K., 1995, ‘‘Nuclear magnetic ordering in silver,’’Ph.D. Thesis, University of Copenhagen, Riso” NationalLaboratory Report R-850(EN).

Lefmann, K., J. T. Tuoriniemi, K. K. Nummila, and A. Metz,1996, ‘‘Neutron thermometry on polarized silver nuclei atsub-microkelvin spin temperatures,’’ Z. Phys. B (in press).

Leggett, A. J., 1974, ‘‘The spin dynamics of an anisotropicFermi superfluid (3He?),’’ Ann. Phys. 85, 11–55.

Leib, J., M. Huebner, S. Gotz, Th. Wagner, and G. Eska, 1995,‘‘NMR-investigations on thallium-samples of high spin-polarization,’’ J. Low Temp. Phys. 101, 253–258.

Lindelof, P. E., I. E. Miller, and G. R. Pickett, 1975, ‘‘Nuclearcooperative ordering in single-crystal praseodymium at lowtemperatures,’’ Phys. Rev. Lett. 35, 1297–1299.

Lindgren, I., and A. Rosen, 1974a, ‘‘Relativistic self-consistent-field calculations with application to atomic hyperfine inter-

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

action. I. Relativistic self-consistent fields. II. Relativistictheory of atomic hyperfine interaction,’’ Case Stud. At. Phys.4, 93–196.

Lindgren, I., and A. Rosen, 1974b, ‘‘Relativistic self-consistent-field calculations with application to atomic hyper-fine interaction. III. Comparison between theoretical and ex-perimental hyperfine-structure results,’’ Case Stud. At. Phys.4, 197–298.

Lindgard, P.-A., 1988a, ‘‘Theory of adiabatic nuclear magneticordering in Cu,’’ Phys. Rev. Lett. 61, 629–632.

Lindgard, P.-A., 1988b, ‘‘Theory of the nuclear magnetic or-dering in Cu in a field,’’ J. Phys. (Paris) Colloq. 49, C8, 2051–2052.

Lindgard, P. A., 1990, ‘‘Theory of a new type of antiferromag-netism in the ideal fcc system of Cu,’’ J. Magn. Magn. Mater.90-91, 138–140.

Lindgard, P. A., 1992, ‘‘Vibrationally reduced magnetic inter-actions in Cu and the magnetic ordering in a magnetic field,’’J. Magn. Magn. Mater. 104–107, 2109–2110.

Lindgard, P. A., X. W. Wang, and B. N. Harmon, 1986, ‘‘Cal-culation of the Ruderman-Kittel interaction and magnetic or-dering in copper,’’ J. Magn. Magn. Mater. 54–57, 1052–1054.

Long, M. W., 1989, ‘‘Effects that can stabilise multiple spin-density waves,’’ J. Phys., Condens. Matter 1, 2857–2874.

Loponen, M. T., R. C. Dynes, V. Narayanamurti, and J. P.Garno, 1981, ‘‘Observation of time-dependent specific heat inamorphous SiO 2,’’ Phys. Rev. Lett. 45, 457–460.

Lounasmaa, O. V., 1974, Experimental Principles and Methodsbelow 1 K (Academic, London).

Lounasmaa, O. V., 1989, ‘‘Nuclear magnetic ordering atnanokelvin temperatures,’’ Phys. Today 42, 26–33.

Lounasmaa, O. V., P. Hakonen, K. Nummila, R. Vuorinen,and J. Martikainen, 1994, ‘‘Negative nanokelvin temperaturesin the nuclear spin systems of silver and rhodium metals,’’Physica B 194–196, 291–292.

Luttinger, J. M., and L. Tisza, 1946, ‘‘Theory of dipole inter-action in crystals,’’ Phys. Rev. 70, 954–964; 72, 257(E).

Marks, J., W. Th. Wenckebach, and N. J. Poulis, 1979, ‘‘Mag-netic ordering of proton spins in Ca(OH)2,’’ Physica B 96,337–340.

Martikainen, J., 1993, ‘‘Nuclear magnetic susceptibility mea-surements in rhodium,’’ Diploma Thesis (in Finnish), Hels-inki University of Technology.

Mattis, D. C., 1985, The Theory of Magnetism, Vol. 2(Springer, Heidelberg).

McEwen, K. A., and W. G. Stirling, 1989, ‘‘Search for nuclearspin polarisation in praseodymium,’’ Physica B 156–157, 754–755.

Meredith, D. J., G. R. Pickett, and O. G. Symko, 1973, ‘‘Ap-plication of a SQUID magnetometer to NMR at low tem-peratures,’’ J. Low Temp. Phys. 13, 607–615.

Merkulov, I. A., Yu. I. Papava, V. V. Ponomarenko, and S. I.Vasiliver, 1988, ‘‘Monte Carlo simulation and theory inGaussian approximation of a phase in the nuclear spin systemof a solid,’’ Can. J. Phys. 66, 135–144.

Miki, T., M. Yanaka, M. Nakagawa, D. Kim, O. Ishikawa, T.Hata, H. Ishii, and T. Kodama, 1992, ‘‘Electrical resistivityanomaly at the nuclear magnetic ordering in a Van Vleckparamagnet PrCu 6,’’ Phys. Rev. Lett. 69, 375–378.

Miller, D. J., and S. J. Frisken, 1988, ‘‘Calculation of theindirect-exchange interaction between nuclear spins in Cuand Ag,’’ J. Appl. Phys. 64, 5630–5632.

Page 134: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

134 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Minor, W., and T. M. Giebultowicz, 1988, ‘‘Studies of fccHeisenberg antiferromagnets by Monte Carlo simulation onlarge spin arrays,’’ J. Phys. (Paris) Colloq. 49, C8, 1551–1552.

Moon, R. M., T. Riste, and W. C. Koehler, 1969, ‘‘Polarizationanalysis of thermal-neutron scattering,’’ Phys. Rev. 181, 920–931.

Moyland, P. L., P. Kumar, J. Xu, and Y. Takano, 1993, ‘‘Cou-pling of the Larmor precession to the correlated motion ofpairs of nuclear spins in noble metals,’’ Phys. Rev. 48, 14020–14022.

Moyland, P. L., T. Lang, J. Xu, E. D. Adams, G. R. Stewart,and Y. Takano, 1995, ‘‘Magnetic susceptibility measurementsin PrBe 13 ,’’ J. Low Temp. Phys. 101, 641–643.

Mueller, R. M., C. Buchal, H. R. Folle, M. Kubota, and F.Pobell, 1980, ‘‘A double-stage nuclear demagnetization re-frigerator,’’ Cryogenics 20, 395–407.

Muething, K., G. G. Ihas, and J. Landau, 1977, ‘‘Metallic ther-mal connectors for use in nuclear refrigeration,’’ Rev. Sci.Instrum. 48, 906–909.

Murao, T., 1971, ‘‘Magnetism of a ‘Nuclei plus electrons’ sys-tem,’’ J. Phys. Soc. Jpn. 31, 683–690.

Murao, T., 1972, ‘‘Magnetism of a ‘Nuclei plus electrons’ sys-tem. II,’’ J. Phys. Soc. Jpn. 33, 33–38.

Murao, T., 1981, ‘‘Magnetism of a nucleus-electron complexsystem in the vicinity of the threshold,’’ J. Phys. Soc. Jpn. 50,3240–3244.

Narath, A., A. T. Fromhold Jr, and E. D. Jones, 1966,‘‘Nuclear spin relaxation in metals: rhodium, palladium, andsilver,’’ Phys. Rev. 144, 428–435.

Nicklow, R. M., R. M. Moon, S. Kawarazaki, N. Kunitomi, H.Suzuki, T. Ohtsuka, and Y. Morii, 1985, ‘‘Nuclear spin order-ing observed by neutron diffraction,’’ J. Appl. Phys. 57, 3784–3788.

Niinikoski, T. O., 1988, ‘‘The principles of detection of darkmatter based on non-resolved fluctuations of excitations insolids,’’ Ann. Phys. (Paris) 13, 143–150.

Niskanen, K. J., L. H. Kjaldman, and J. Kurkijarvi, 1982, ‘‘Or-dered phases of the combinated Ruderman-Kittel and dipole-dipole interaction on different lattices and their stability,’’ J.Low Temp. Phys. 49, 241–249.

Niskanen, K. J., and J. Kurkijarvi, 1981, ‘‘The transition tem-perature of Cu nuclei to an ordered state,’’ J. Phys. C 14,5517–5522.

Niskanen, K. J., and J. Kurkijarvi, 1983, ‘‘Linked cluster ex-pansion for quantum spins,’’ J. Phys. A 16, 1491–1504.

Nummila, K. K., J. T. Tuoriniemi, R. T. Vuorinen, K. Lef-mann, A. Metz, and F. B. Rasmussen, 1997, ‘‘Neutron diffrac-tion studies of nuclear magnetic ordering in silver,’’ Phys.Rev. B (to be published).

Oguchi, T., 1960, ‘‘Theory of spin-wave interactions in ferro-and antiferromagnetism,’’ Phys. Rev. 117, 117–123.

Oguchi, T., H. Nishimori, and Y. Taguchi, 1985, ‘‘The spinwave theory in antiferromagnetic Heisenberg model on facecentered cubic lattice,’’ J. Phys. Soc. Jpn. 54, 4494–4497.

Oh, G.-H., Y. Ishimoto, T. Kawae, M. Nakagawa, O. Ishikawa,T. Hata, and T. Kodama, 1994, ‘‘Cooling of 3He- 4He dilutesolution down to 97 mK. Thermal boundary resistance be-tween dilute solution and metal powder,’’ J. Low Temp. Phys.95, 525–546.

Oja, A. S., 1984, ‘‘Magnetic ordering of copper nuclei,’’ Di-ploma Thesis (in Finnish), Helsinki University of Technol-ogy.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Oja, A. S., 1987, ‘‘Nuclear magnetism in metals,’’ Phys. Scr.T19, 462–468.

Oja, A. S., 1991, ‘‘Nuclear magnetic ordering in copper andsilver at nanokelvin temperatures,’’ Physica B 169, 306–315.

Oja, A. S., A. J. Annila, and Y. Takano, 1988, ‘‘Resplitting ofexchange-merged NMR absorption lines at high spin polar-izations,’’ Phys. Rev. B 38, 8602–8608.

Oja, A. S., A. J. Annila, and Y. Takano, 1990, ‘‘Cross relax-ation by single spin flips in silver at nanokelvin tempera-tures,’’ Phys. Rev. Lett. 65, 1921–1924.

Oja, A. S., A. J. Annila, and Y. Takano, 1991, ‘‘Investigationsof nuclear magnetism in silver down to picokelvin tempera-tures. I,’’ J. Low Temp. Phys. 85, 1–24.

Oja, A. S., and P. Kumar, 1984, ‘‘Helical structures of nuclearspins in a metal,’’ Physica B 126, 451–452.

Oja, A. S., and P. Kumar, 1987, ‘‘Indirect nuclear spin interac-tions and nuclear ordering in metals,’’ J. Low Temp. Phys. 66,155–167.

Oja, A. S., and H. E. Viertio, 1987, ‘‘Magnetically orderedstructures of nuclear spins in fcc metals,’’ Jpn. J. Appl. Phys.26-3, 441–442.

Oja, A. S., and H. E. Viertio, 1992, ‘‘Nuclear magnetic order-ing in copper: the spin structure in the high-field phase atBi[111],’’ J. Magn. Magn. Mater. 104–107, 908– 910.

Oja, A. S., and H. E. Viertio, 1993, ‘‘Antiferromagnets withanisotropic spin-spin interactions: Stability of the zero-fieldstructure in an external field,’’ Phys. Rev. B 47, 237–253.

Oja, A. S., X.-W. Wang, and B. N. Harmon, 1989, ‘‘First-principles study of the conduction-electron-mediated interac-tions between nuclear spins in copper metal,’’ Phys. Rev. B39, 4009–4021.

Okamoto, T., H. Fukuyama, H. Akimoto, H. Ishimoto, and S.Ogawa, 1994, ‘‘Direct demagnetization cooling of high-density solid 3He,’’ Phys. Rev. Lett. 72, 868–871.

Osheroff, D. D., M. C. Cross, and D. S. Fisher, 1980, ‘‘Nuclearantiferromagnetic resonance in solid 3He,’’ Phys. Rev. Lett.44, 792–795.

Pershan, P. S., 1960, ‘‘Cross relaxation in LiF,’’ Phys. Rev. 117,109–116.

Pickett, G. R., 1988, ‘‘Microkelvin physics,’’ Rep. Prog. Phys.51, 1295–1340.

Pirnie, K., P. J. Wood, and J. Eve, 1966, ‘‘On high temperaturesusceptibilities of Heisenberg model ferromagnetics and anti-ferromagnetics,’’ Mol. Phys. 11, 551–577.

Pobell, F., 1982, ‘‘The quest for ultralow temperatures: Whatare the limitations?’’ Physica B 109-110, 1485–1498.

Pobell, F., 1988, ‘‘La quete du zero absolu,’’ Recherche 19,784–789.

Pobell, F., 1992a, Matter and Methods at Low Temperatures(Springer, Berlin).

Pobell, F., 1992b, ‘‘Nuclear refrigeration and thermometry atmicrokelvin temperatures,’’ J. Low Temp. Phys. 87, 635–649.

Pobell, F., 1994, ‘‘Nuclear magnetic ordering in metals,’’Physica B 197, 115–123.

Poitrenaud, J., 1967, ‘‘Resonance magnetique nucleaire dans lerubidium et le cesium metalliques,’’ J. Phys. Chem. Solids 28,161–170.

Poitrenaud, J., and J. M. Winter, 1964, ‘‘Resonance nucleairedans l’argent metallique,’’ J. Phys. Chem. Solids 25, 123– 127.

Pollack, L., E. N. Smith, J. M. Parpia, and R. C. Richardson,1992, ‘‘Novel low-temperature cross relaxation in nuclearquadrupole resonance,’’ Phys. Rev. Lett. 69, 3835–3838.

Page 135: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

135A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Price, D. L., and K. Skold, 1986, ‘‘Introduction to neutron scat-tering,’’ in Neutron Scattering: Methods of ExperimentalPhysics, edited by K. Skold and D. L. Price (Academic, Lon-don), Vol. 23, Part A, pp. 1–98.

Provotorov, B. N., 1962, ‘‘Quantum statistical theory of crossrelaxation,’’ Zh. Eksp. Teor. Fiz. 42, 882–888 [Sov. Phys.JETP 15, 611–614 (1962)].

Purcell, E. M., and R. V. Pound, 1951, ‘‘A nuclear spin systemat negative temperature,’’ Phys. Rev. 81, 279–280.

Ramakrishnan, S., and G. Chandra, 1993, ‘‘Nuclear magnetismin metals and alloys,’’ in Selected Topics in Magnetism, Fron-tiers in Solid State Sciences, No. 2, edited by L. C. Gupta andM. S. Multani (World Scientific, Singapore), pp. 409–441.

Ramsey, N. F., 1956, ‘‘Thermodynamics and statistical me-chanics at negative absolute temperatures,’’ Phys. Rev. 103,20–28.

Richardson, R. C., and E. N. Smith, 1988, Experimental Tech-niques in Condensed Matter Physics at Low Temperatures(Addison Wesley, New York).

Rivier, N., and V. Zlatic, 1972, ‘‘Temperature dependence ofthe resistivity due to localized spin fluctuations (Part I),’’ J.Phys. F 2, L87–L92; 2, ‘‘Temperature dependence of the re-sistivity due to localized spin fluctuations II–Coles alloys,’’L99–L104.

Rodak, M. I., 1971, ‘‘Behavior of a system of spins close infrequency and coupled by cross relaxation,’’ Zh. Eksp. Teor.Fiz. 61, 832–842 [Sov. Phys. JETP 34, 443–448 (1972)].

Roger, M., J. H. Hetherington, and J. M. Delrieu, 1983, ‘‘Mag-netism in solid 3He,’’ Rev. Mod. Phys. 55, 1–64.

Roinel, Y., G. L. Bachella, O. Avenel, V. Bouffard, M. Pinot,P. Roubeau, P. Meriel, and M. Goldman, 1980, ‘‘Neutron dif-fraction study of nuclear magnetic ordered phases and do-mains in lithium hydride,’’ J. Phys. (Paris) Lett. 41, 123–125.

Roinel, Y., V. Bouffard, G. L. Bacchella, M. Pinot, P. Meriel,P. Roubeau, O. Avenel, M. Goldman, and A. Abragam, 1978,‘‘First study of nuclear antiferromagnetism by neutron dif-fraction,’’ Phys. Rev. Lett. 41, 1572–1574.

Roinel, Y., V. Bouffard, C. Fermon, M. Pinot, F. Vigneron, G.Fournier, and M. Goldman, 1987, ‘‘Phase diagrams of or-dered nuclear spins in LiH: a new phase at positive tempera-ture?’’ J. Phys. (Paris) 48, 837–845.

Roser, M. R., and L. R. Corruccini, 1990, ‘‘Dipolar ferromag-netic order in a cubic system,’’ Phys. Rev. Lett. 65, 1064–1067.

Roshen, W. A., and W. F. Saam, 1980, ‘‘Effect of Kondo im-purities on the Korringa relaxation time,’’ Phys. Rev. B 22,5495–5500.

Roshen, W. A., and W. F. Saam, 1982, ‘‘More exact field de-pendence of nuclear spin relaxation of the host metal in theKondo systems,’’ Phys. Rev. B 26, 2644–2647.

Rossat-Mignod, J., P. Burlett, O. Vogt, and G. H. Lander,1979, ‘‘Magnetization and neutron scattering studies of mul-tiaxial magnetic ordering in USb0.9Te 0.1 ,’’ J. Phys. C 12,1101–1112.

Rowland, T. J., 1960, ‘‘Nuclear magnetic resonance in copperalloys. Electron distribution around solute atoms,’’ Phys.Rev. 119, 900–912.

Ruderman, M. A., and C. Kittel, 1954, ‘‘Indirect exchange cou-pling of nuclear magnetic moments by conduction electrons,’’Phys. Rev. 96, 99–102.

Rushbrooke, G. S., G. A. Baker, and P. J. Wood, 1974,‘‘Heisenberg model,’’ in Phase Transitions and Critical Phe-nomena, edited by C. Domb and M. S. Green (Academic,New York), Vol. 3, pp. 245–356.

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Ryhanen, T., H. Seppa, R. Ilmoniemi, and J. Knuutila, 1989,‘‘SQUID magnetometers for low-frequency applications,’’ J.Low Temp. Phys. 76, 287–386.

Schermer, R. I., and M. Blume, 1968, ‘‘Polarization effects inslow-neutron scattering. III. Nuclear polarization,’’ Phys.Rev. 166, 554–561.

Schmid, D., 1973, ‘‘Nuclear magnetic double resonance–Principles and applications in solid state physics,’’ in SpringerTracts of Modern Physics 68 (Springer, Berlin/Heidelberg),pp. 1–75.

Schroder-Smeibidl, B., P. Smeibidl, G. Eska, and F. Pobell,1991, ‘‘Nuclear specific heat of thallium,’’ J. Low Temp. Phys.85, 311–320.

Schwark, M., F. Pobell, W. P. Halperin, C. Buchal, J. Hanssen,M. Kubota, and R. M. Muller, 1983, ‘‘Ortho-para conversionof hydrogen in copper as origin of time dependent heatleaks,’’ J. Low Temp. Phys. 53, 685–694.

Shibata, F., and Y. Hamano, 1982, ‘‘Spin relaxation in metalsat very low temperatures,’’ Solid State Commun. 44, 921–925.

Shigematsu, T., K. Morita, Y. Fujii, T. Shigi, M. Nakamura,and M. Yamaguchi, 1992, ‘‘Investigation of annealing effectsof ultra pure copper,’’ Cryogenics 32, 913–915.

Siemensmeyer, K., K. Kakurai, M. Steiner, T. A. Jyrkkio, M.T. Huiku, and K. N. Clausen, 1990, ‘‘Neutron scattering in-vestigation of the ordered state of the nuclear spins in Cu at60 nK,’’ J. Appl. Phys. 67, 5433–5435.

Siemensmeyer, K., and M. Steiner, 1992, ‘‘Magnetostrictiveand quadrupolar anisotropy in nuclear magnetic fcc systems,’’Z. Phys. B 89, 305–311.

Slichter, C. P., 1990, Principles of Magnetic Resonance(Springer-Verlag, Berlin), pp. 219–246.

Smart, J. B., 1966, Effective Field Theories of Magnetism(Saunders, Philadelphia).

Smeibidl, P., B. Schroder-Smeibidl, and F. Pobell, 1994, ‘‘Spinlattice relaxation measurements on Cu at very low tempera-tures,’’ Physica B 194–196, 337–338.

Smythe, W. R., 1950, Static and Dynamic Electricity (McGraw-Hill, New York).

Soini, J. K., 1982, ‘‘Further studies of nuclear ordering in me-tallic copper,’’ Helsinki University of Technology ReportTKK-F-A471.

Soulen, R. J., and R. B. Dove, 1979, Standard Reference Mate-rial 768, Superconductive Thermometric Fixed Point Device(0.015 K–0.208 K), US National Bureau of Standards.

Steiner, M., 1985, ‘‘Neutron scattering studies of nuclear ori-entation in the mK range,’’ in Neutron Scattering in the Nine-ties (IAEA report), pp. 185–190.

Steiner, M., 1990, ‘‘The use of the spin-dependent scatteringlength: New areas and improved techniques for neutron-scattering,’’ Phys. Scr. 42, 367–384.

Steiner, M., 1993, ‘‘Nuclear magnetic order: Use of spin depen-dent amplitudes,’’ Internat. J. Mod. Phys. B 7, 2909–2944.

Steiner, M., L. Bevaart, Y. Ajiro, A. J. Millhouse, K. Ohlhoff,G. Rahn, H. Dachs, U. Scheer, and B. Wanklyn, 1981, ‘‘Ob-servation of hyperfine-enhanced nuclear polarisation inCoF2 by means of neutron diffraction,’’ J. Phys. C 14, L597–L602.

Steiner, M., A. Metz, K. Siemensmeyer, O. V. Lounasmaa, J.T. Tuorinemi, K. K. Nummila, R. T. Vuorinen, K. N.Clausen, K. Lefmann, and F. B. Rasmussen, 1996, ‘‘Neutrondiffraction determination of the nuclear spin ordering in Cuand Ag at nano- and subnano-K temperatures,’’ J. Appl.Phys. 79, 5078–5080.

Page 136: Nuclear magnetic ordering in simple metals at positive and …yast/Articles/MyArt/Ordering.pdf · 2004-01-02 · Nuclear magnetic ordering in simple metals at positive and negative

136 A. S. Oja and O. V. Lounasmaa: Nuclear magnetic ordering in simple metals

Steiner, M., K. Siemensmeyer, K. D. Ohlhoff, G. Rahn, M.Kubota, and S. M. Smith, 1986, ‘‘A polarized neutron diffrac-tion study of nuclear polarization and order in HoVo4,’’ J.Magn. Magn. Mater. 54–57, 1333–1334.

Suzuki, H., Y. Koike, Y. Karaki, M. Kubota, and H. Ishimoto,1994, ‘‘Nuclear spin order in Sc metal,’’ Physica B 194–196,249–250.

Suzuki, T., H. Kondo, H. Yano, S. Abe, Y. Miura, and T.Mamiya, 1991, ‘‘Zeeman-exchange relaxation time of h.c.p.solid 3He near the nuclear ordering temperature,’’ Europhys.Lett. 16, 467– 471.

Symko, O. G., 1969, ‘‘Nuclear cooling using copper and in-dium,’’ J. Low Temp. Phys. 1, 451–467.

Takano, Y., N. Nishida, Y. Miura, H. Fukuyama, H. Ishimoto,S. Ogawa, T. Hata, and T. Shigi, 1985, ‘‘Magnetization of hcpsolid 3He,’’ Phys. Rev. Lett. 55, 1490–1493.

Tuoriniemi, J. T., 1995, ‘‘Neutron Experiments on NuclearMagnetism in Copper and Silver,’’ Ph.D. Thesis (HelsinkiUniversity of Technology).

Tuoriniemi, J. T., K. Lefmann, K. K. Nummila, and A. Metz,1995, ‘‘Neutron studies of highly polarized silver nuclei atnanokelvin spin temperatures,’’ Helsinki University of Tech-nology Report TKK-F-A744.

Tuoriniemi, J. T., K. K. Nummila, K. Lefmann, R. T. Vuo-rinen, and A. Metz, 1996, ‘‘Neutron thermometry applied tomagnetization and spin-lattice relaxation measurements onsilver nuclei,’’ Z. Phys. B (to be published).

Tuoriniemi, J. T., K. K. Nummila, R. T. Vuorinen, O. V. Lou-nasmaa, A. Metz, K. Siemensmeyer, M. Steiner, K. Lefmann,K. N. Clausen, and F. B. Rasmussen, 1995, ‘‘Neutron experi-ments on antiferromagnetic nuclear order in silver at pi-cokelvin temperatures,’’ Phys. Rev. Lett. 75, 3744–3747.

Urbina, C., J. F. Jacquinot, and M. Goldman, 1982, ‘‘Rotatingtransverse helical nuclear ordering,’’ Phys. Rev. Lett. 48, 206–209.

Urbina, C., J. F. Jacquinot, and M. Goldman, 1986, ‘‘Rotatingtransverse nuclear helimagnetism in CaF2: I. Prediction andexperimental study,’’ J. Phys. C 19, 2275–2297.

Van der Zon, C. M., G. D. Van Velzen, and W. Th. Wenck-ebach, 1990, ‘‘Nuclear magnetic ordering in Ca(OH) 2. III.Experimental determination of the critical temperature,’’ J.Phys. (Paris) 51, 1479–1488.

Van Kesteren, H. W., W. Th. Wenckebach, and J. Schmidt,1985, ‘‘Production of high, long-lasting, dynamic proton po-larization by way of photoexcited triplet states,’’ Phys. Rev.Lett. 55, 1642–1644.

Van Vleck, J. H., 1937, ‘‘The influence of dipole-dipole cou-pling on the specific heat and susceptibility of a paramagneticsalt,’’ J. Chem. Phys. 5, 320–337.

Van Vleck, J. H., 1948, ‘‘The dipolar broadening of magneticresonance lines in crystals,’’ Phys. Rev. 74, 1168–1183.

Van Vleck, J. H., 1957, ‘‘The concept of temperature in mag-netism,’’ Nuovo Cimento Suppl. 6, 1081–1100.

Veuro, M. C., 1978, ‘‘A double bundle nuclear demagnetiza-tion refrigerator for cooling 3He,’’ Acta Polytech. Scand. Ph.122.

Viertio, H. E., 1990, ‘‘Thermally stabilized stuctures of thenuclear spin system in silver,’’ Phys. Scr. T33, 168–175.

Viertio, H., 1992, ‘‘Theoretical studies on nuclear magnetic or-dering in copper and silver,’’ Ph.D. thesis (Helsinki Univer-sity of Technology).

Rev. Mod. Phys., Vol. 69, No. 1, January 1997

Viertio, H. E., and A. S. Oja, 1987, ‘‘Ground state of nuclearspins in fcc metals,’’ Phys. Rev. B 36, 3805–3808.

Viertio, H. E., and A. S. Oja, 1989, ‘‘Nuclear magnetism incopper at nanokelvin temperatures,’’ in Symposium on Quan-tum Fluids and Solids 21989, AIP Conference ProceedingsNo. 194, edited by G. G. Ihas and Y. Takano (AIP, NewYork), pp. 305–315.

Viertio, H. E., and A. S. Oja, 1990a, ‘‘Nuclear antiferromag-netism in copper: Interplay of (0, 2

3, 23) and (1,0,0) order,’’ Phys.

Rev. B 42, 6857–6860.Viertio, H. E., and A. S. Oja, 1990b, ‘‘Nuclear antiferromag-

netism in copper: Predictions for spin structures in a magneticfield,’’ Physica B 165–166, 799–800.

Viertio, H. E., and A. S. Oja, 1992, ‘‘Mean-field calculationand Monte Carlo simulation of ferromagnetic ordering atnegative temperatures,’’ J. Magn. Magn. Mater. 104–107,915–917.

Viertio, H. E., and A. S. Oja, 1993, ‘‘Interplay of three antifer-romagnetic modulations in the nuclear-spin system of cop-per,’’ Phys. Rev. B 48, 1062–1076.

Villain, J., 1959, ‘‘La structure des substances magnetiques,’’ J.Phys. Chem. Solids 11, 303–309.

Villain, J., R. Bidaux, J.-P. Carton, and R. Conte, 1980, ‘‘Orderas an effect of disorder,’’ J. Phys. (Paris) 41, 1263–1272.

Vuorinen, R., 1992, ‘‘Ferromagnetic ordering of silver nuclei atnegative absolute temperatures,’’ Diploma Thesis (in Finn-ish) (Helsinki University of Technology).

Vuorinen, R. T., P. J. Hakonen, W. Yao, and O. V. Lounas-maa, 1995, ‘‘Susceptibility and relaxation measurements onrhodium metal at positive and negative spin temperatures inthe nanokelvin range,’’ J. Low Temp. Phys. 98, 449–487.

Webb, R. A., 1977, ‘‘New technique for improved low-temperature SQUID NMR measurements,’’ Rev. Sci. In-strum. 48, 1585–1594.

White, S. J., M. R. Roser, J. Xu, J. T. Van der Noordaa, and L.R. Corruccini, 1993, ‘‘Dipolar magnetic order with largequantum spin fluctuations in a diamond lattice,’’ Phys. Rev.Lett. 71, 3553–3556.

Windsor, C. G., 1986, ‘‘Experimental techniques,’’ in NeutronScattering: Methods of Experimental Physics, edited by K.Skold and D. L. Price (Academic, London), Vol. 23, Part A,pp. 197–258.

Wolf, D., 1979, Spin Temperature and Nuclear-Spin Relaxationin Matter (Clarendon, Oxford).

Wysin, G. M., and A. R. Bishop, 1990, ‘‘Dynamic correlationsin a classical two-dimensional Heisenberg antiferromagnet,’’Phys. Rev. B 42, 810–819.

Yano, H., H. Kondo, T. Suzuki, Y. Minamide, T. Kato, Y.Miura, and T. Mamiya, 1990, ‘‘Magnetization and ac suscep-tibility of hcp solid 3He near the nuclear ordering tempera-ture,’’ Phys. Rev. Lett. 65, 3401–3404.

Yin, S., and P. J. Hakonen, 1991, ‘‘Electron-beam welded Cu-to-Ag joints for thermal contact at low temperatures,’’ Rev.Sci. Instrum. 62, 1370–1371.

Yosida, K., 1957, ‘‘Magnetic properties of Cu-Mn alloys,’’Phys. Rev. 106, 893–898.

Zachariassen, W. H., 1967, ‘‘A general theory of X-ray diffrac-tion in crystals,’’ Acta Crystallogr. 23, 558–564.

Zimmermann, J., and G. Weber, 1981, ‘‘Thermal relaxation oflow-energy excitations in vitreous silica,’’ Phys. Rev. Lett. 46,661–664.