• Aucun résultat trouvé

ELECTRONIC STRUCTURE OF THE LIGHT ACTINIDES

N/A
N/A
Protected

Academic year: 2021

Partager "ELECTRONIC STRUCTURE OF THE LIGHT ACTINIDES"

Copied!
11
0
0

Texte intégral

(1)

HAL Id: jpa-00216775

https://hal.archives-ouvertes.fr/jpa-00216775

Submitted on 1 Jan 1976

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

ELECTRONIC STRUCTURE OF THE LIGHT

ACTINIDES

B. Dunlap

To cite this version:

(2)

ELECTRONIC STRUCTURE IN METALS AND ALLOYS.

ELECTRONIC STRUCTURE OF THE LIGHT ACTINIDES

B. D. DUNLAP

Argonne National Laboratory, Argonne, Illinois 60439, U. S. A.

R6sum6. - Notre comprehension des propriCt6s des actinides legers a largement progresse au cours des dernikres annks. Bien que ces elements appartenant la sQie de transition 5f soient formellement Bquivalents h ceux de la skrie des lanthanides (4f), leur comportement est beaucoup plus varie du fait de I'extension radiale plus &levee et de la facilite d'ionisation de I'orbitale 5f. Cette revue dBcrit quelques thkmes #actualit&, en particulier ceux ou la contribution des techniques #interactions hyperfines a et6 importante. I1 s'agit d'etudes de divers phenomknes magnetiques, de la systematique des deplacements isom6riques ainsi que d'effets de relaxation paramagnetique.

Abstract. - In the last few years, considerable advances have been made in our understanding of the properties of the light actinides. Although these 5f transition elements are formally equivalent to the lanthanide (4f) elements, they show a much more varied behaviour due to the larger spatial extent and ionizability of the 5f electrons. A review is given of some areas of current interest, especially where hyperfine techniques have played an active role. It includes studies of a variety of magnetic phenomena, systematics of isomer shift measurements, and studies of paramagnetic relaxation.

1. Introduction. - Mossbauer effect measurements in the actinide series were first obtained in 1964 using the a decay of 241Am and the

p

decay of 237U as sources for the 60 keV y-ray of 237Np [l]. At that time, our understanding of the electronic properties of the actinides was at a very early stage. Subsequently, a large number of theoretical and experimental approa- ches have been brought to bear on the properties of these materials. As a result, substantial progress has been made in obtaining some overall views on the ways these materials function, and how they differ from materials comprised of elements in other parts of the periodic table [2]. In this process, the Mossbauer effect measurements have proven to be valuable, especially when considered together with results obtained from other techniques. The special ability of hyperfine measurements to provide information on local magnetic moments, local electronic densities, time dependent phenomena, etc. has provided both a source of original data and problems, as well as a valuable supplement to data which emphasizes the bulk properties of the materials.

At present, Mossbauer effect measurements have been performed in some isotope of all elements in the first half of the actinide series [3]. Of these, however, several have severe limitations. The most common occurrence is a transition between two nuclear levels in a rotational band. These transitions are characterized by : (i) almost unmeasurable isomer shifts, (ii) a short lifetime (- 0.1 ns) resulting in wide lines. and (iii) severe internal conversion (a several hundred). The most usable cases are the even isotopes of ura- nium, where some limited measurements have been

made [4, 51 and for some problems these resonances will continue to be useful. However, the availability of two excellent resonances has almost completely overshadowed the low resolution isotopes. These are the 59.6 keV,

3

-+

3

transition in 237Np [6, 71 and the 83.9 keV,

3

-t

3

transition in 243Am [8, 91. Both of

these combine large isomer shifts and excellent resolu- tion with relative ease of measurement. The 243Am resonance, however, has the difficulty of having only a 5 hour source activity in the

p

decay of 2 4 3 P ~ . While thk source is easily obtained if one is near a reactor, it is clearly difficult otherwise. In addition, there is a tendency for Am materials to prefer the trivalent state which, in analogy to its lanthanide analogue, Eu3+, shows no hyperfine splittings. On the contrary, 237Np most commonly uses as source the 458 yr a-decay of

241Am. Also, Np ions occur in several charge states, all of which show marked hyperfine interactions. Consequently, the literature has been dominated in the past by work with 237Np, and will probably continue to be so in the future, with a smaller effort on 243Am systems and still less from the other isotopes, primarily uranium.

No attempt will be made here to survey thoroughly the actinide literature, since such a survey is available of the earlier work 131. Rather the aim here is to show some of the more interesting aspects of actinide beha- viour and try to convince the reader that this is an area where much interesting physics occurs of signifi- cance on a scale broader than might first appear. In the following we will first discuss briefly the basic features which make the electronic properties of the actinides unusual, and will then consider a number of

(3)

C6-308 B. D. DUNLAP

problems where hyperfine interaction measurements currently are active or are likely to become increasingly active in the near future.

2. Properties of 5f electrons.

-

The actinides form a transition series of elements in which the 5f shell is being filled. They are thus formally analogous in the periodic table to the lanthanides. However, the first thing to understand is that 5f electrons differ in impor- tant aspects from 4f electrons, so actinides should not be thought of as being heavy lanthanides. A conside- ration of atomic calculations for binding energies and wavefunctions in the two systems show two primary differences [l01 related to the 5f electron binding energy and spatial distribution.

In the lanthanide series, the 4f electrons are strongly bound compared to the outer 5d and 6s electrons. When these atoms are placed in a solid material, the 5d and 6s electrons and sometimes one 4f electron are ionized, leaving a very stable trivalent ion 1111. In the actinides, the 5f electrons are comparable in energy with the 6d and 7s electrons, and hence can be easily ionized. This simple fact causes many effects. In salts, the actinide ions can be formed in several charge states. For example, neptunium (atomic configuration 5f4 6d7s2) forms salts with ionic states varying from Np3+ to Np7'. In metals, the 5f electrons are expected to contribute to the conduction processes along with the 6d and 7s electrons. While one may assume that lanthanides are in a trivalent state even in metallic systems, it is very difficult to make an a priori

guess for the nominal valence of an actinide ion in a metallic system. Indeed, one may argue that such a concept is not even well-defined, similar to the diffi- culties involved in treating d transition element metals. Spatially, the 5f electrons are considerably more extended than 4f electrons. While this does not appear at first glance to be a profound statement in view of the fact that the higher quantum number orbitals must lie farther from the nucleus, it is nonetheless a fact of great importance. Examinations of the spatial distribu- tion of 5f electrons obtained from relativistic self- consistent-field atomic calculations show [l01 that a significant electron density is still present at distances corresponding to typical half-interatomic spacings in solids. In salts, this means that the 5f electrons enter into the bonding process, making the chemistry consi- derably more complex than with the lanthanides. In addition, crystal field interactions are larger than in the lanthanides. In metals, the overlap of adjacent 5f orbitals causes these electrons to de-localize into conduction bands, the width of the band and the extent of the de-localization being dependent on the actinide-actinide spacing. This simple concept has been used by Hill [l21 to classify the magnetic proper- ties of metallic systems solely on the basis of the intera- tomic spacing. When the actinide ions are near one another, as in the elemental metals, the 5f electrons are

in bands sufficiently wide to inhibit magnetism, and the light actinide metals are observed to be Pauli parama- gnets. In intermetallics, where the spacing between the actinides is larger, more localized behaviour exists and varied magnetic phenomena are seen. Hill, in fact, suggests a critical spacing below which ordered magne- tism can not occur. While such arguments are undoub- tedly over-simplified, they certainly provide an impor- tant framework in which to consider the magnetic phenomena. As one goes to the heavier actinides, there is a spatial contraction of the 5f electronic distribution similar to the well-known lanthanide contraction, and the behaviour becomes more loca- lized even for the simple metals. Eventually, beginning roughly at Am, behaviour more like the lanthanides is expected.

These two basic features, i. e. the smaller ionization energies of the 5f electrons and their larger spatial distribution, provide the essential first-order informa- tion that must be kept in mind as the actinides are considered. It is clear, however, that these features make these materials rather complex in which common approximations (like the use of purely atomic 4f orbitals for lanthanides ions in solids) can not be trivially applied. Hence much of the work to present has consisted simply of the collection of data to provide a base for evaluating systematic behaviour. At present, some trends are visible, and some direction for future work can be seen. We will now consider some of these areas.

(4)

ELECTRONIC STRUCTURE OF THE LIGHT ACTINIDES C6-309

or even paramagnetic behaviour, presumably because of a tendency to form the trivalent state. Like its lanthanide counterpart, Eu3+, this is a configuration with no magnetic moment. (3) There are classes of compounds where magnetic ordering occurs, but the 5f electrons do not have localized character. These are characterized by a strong dependence of the magnetic properties on lattice parameter, small hyperfine fields, non-Curie-Weiss susceptibility, low magnetic entropy in the specific heat measurements and a large electro- nic specific heat due to f-electron contributions. For the present discussion, we confine ourselves to neptunium systems, where the most roundly deve- loped data is available. In keeping with the above, N p metal shows no magnetic ordering and Mossbauer effect measurements on Np metal in an external field indicate that no paramagnetic moment is present on the Np ion 1141.

In intermetallics, magnetic ordering is frequently observed. In a few cases, a systematic study has been made of several compounds of similar structure, and magnetic variety rather like what is observed in the lanthanides has been noted. For example, in the sodium chloride compounds NpM (M = N, P, As, Sb, S) many kinds of magnetism are observed. Ferromagne- tism is seen in NpN, antiferromagnetism with multiple transitions in NpN and NpAs, simple antiferromagne- tism in NpSb [15], and antiferromagnetism with a first-order phase transition in NpS [16]. Similarly in the compounds NpM,Si, (M = Cr, Mn, Fe, CO, Ni,

Cu) many kinds of effects are present [17]. NpMn,Si, orders ferromagnetically with a transition temperature above room temperature, NpCu,Si, has a 34 K

transition which is also first order ; and the others show complex antiferromagnetism with transition

temperatures ranging between 33 and 87 K. Some dis- cussion of these results in terms of crystal field inter- actions based on purely ionic states has been given [l71 and it is possible to understand some of the general features observed. In general, one may anticipate that systems such as these will yield information much of the kind that has been obtained in lanthanide inter- metallics concerning exchange interactions, crystal fields, etc.

It is interesting to compare observed hyperhe fields, H,,, with the electronic magnetic moment, p, in the magnetically ordered materials [18]. The hyperfine field can be written as the sum of an orbital and

a

core- polarization term. If we consider a single manifold in the limit of L-S coupling, then

where g, is the Land6 factor,

<

J

II

N

11

J

I1

>

is a reduced matrix element dependent on configuration,

<

r - 3

>

is the radial parameter for the f electrons and

X

is approximately a constant. Now if

<

r - 3

>

is a constant at the free-ion value for a number of different

materials, then a plot of experimental values of H,, vs. p will be a straight line extrapolating through the free-ion point. On the other hand, if the electrons are de-localized for some materials, then

<

r - 3

>

will be altered and the simple relationship will fail. Hyperfine fields obtained by the Mossbauer effect and magnetic moments obtained by neutron diffraction [l91 are shown in figure 1, as well as calculated free-ion values.

FIG. 1. - Magnetic hyperfine field Hhf as determined by the

Mossbauer effect vs. magnetic moment p as determined by

neutron diffraction. Experimental points are given by closed

circles and calculated free-ion values by open circles.

As one sees, for a great many compounds in various crystal structures, a localized description seems to be appropriate. It may be possible that a narrow band description can give values of

<

r - 3

>

that are not severely different from the free-ion values, but to present no attempt has been made to show this.

Given such a strong correlation as figure 1, the exceptions become especially interesting. The most thoroughly examined case is NpOs,. A good deal of work has been done on the neptunium Laves phase systems because they appear to form a bridge between the localized and itinerant systems. The variation in lattice parameter a, is sufficient to significantly affect the magnetic character of different alloys in the series [20-231. The compound with the largest lattice spacing, NpAI, with a, = 7.785

A,

is a ferromagnet with

T,

= 56 K and a well-defined Curie-Weiss behaviour in the susceptibility characteristic of localized magnetic moments. The smallest lattice spacing (except for compounds with magnetic 3d transition elements) occurs for NpRu, (a, = 7.446) which shows tempera- ture independent paramagnetism with no magnetic ordering, characteristic of a Pauli paramagnetic. NpOs, is intermediate (a, = 7.528

A)

and undergoes a ferromagnetic transition with Tc 9 K. It shows some characteristics of localized behaviour in that the 5f electrons still produce a substantial hyperfine field. However, it is clear that a purely localized approach will not be sufficient to describe the material as evi- denced by the following : (i) The susceptibility above

(5)

measurements show a magnetic entropy of only 0.2 R (In 2) [24]. (iii) Even in the saturated magnetic state well below the magnetic transition temperature, application of an external field of 40 kOe causes an increase in the hyperfine field and magnetic moment of about 15

%

[25].

It is of interest to compare the relative values of the magnetic moment as deduced by different measure- ments for these materials [25]. In NpAl,, an apparently well localized case, the neutron diffraction and hyper- fine measurements both indicate a moment of 1.5 ,UB,

while the high field bulk magnetization data give 1.2 pB.

This difference is common in actinides. It is attributed to conduction electron polarisation effects which are measured in the bulk magnetization, but do not significantly affect the more local measurements of hyperfine fields or neutron form factors. On the other hand, in NpOs, this ordering is reversed, with both bulk magnetization and hyperfine fields (using Fig. 1) giving a moment of 0.4 p, while the neutron diffraction result is lower, 0.25 p,. This is due to the fact that the neutron form factor is more sensitive to the outer portions of the electronic distribution whereas the hyperfine field, depending on

<

r - 3

>,

is more

affected by the inner portions. This is illustrated more fully in figure 2. In the upper portion of the figure is shown the radial charge density for a Np3+ ion as obtained by a relativistic (Slater-Fock) atomic calcula- tion [26]. In figure 2(b), we have plotted (along with some other things) the quantity

which shows the way in which the total

<

r - 3

>

determining the hyperfine field is built up as different portions of the electronic cloud are added to the inte- gral. One sees that electrons within a radius of about 1

A

contribute more than 90

%

of the total

<

r-,

>.

A similar calculation is shown in figure 2c for the neutron from factor and one sees that electrons out to 2 or 3

A

still contribute to the neutron measurement. When the electrons begin to de-localize from purely atomic orbitals, the outer parts of the distribution will presumably be affected first. This will cause a pronounced change in the neutron form factor and a weaker effect on

<

r - 3

>.

A data analysis which

assumes a free-ion form factor, or a free-ion

<

r - 3

>

as in figure 1, will produce unusual moments and can lead to the change in ordering observed for NpOs,.

Another interesting case is NpSn,. Neutron diffrac- tion data is not available. However, specific heat [27], magnetic susceptibility [28], and Mossbauer effect [29] measurements all show a magnetic transition at T

-

10 K. The susceptibility shows only a weak ordering, but appears to indicate antiferromagnetism. The anti-ferromagnetic behaviour is more clearly seen in Mossbauer effect measurements at the Sn site. Two magnetically inequivalent sites are seen withlone

( c )

-

-

-

I 2 3 4 r (atomic units)

FIG. 2.

-

Plot of radial charge distribution for 5f electrons

in Np3+, and the fractional contribution of different portions of this distribution to various parameters (see text).

having a transferred hyperfine field and the other having none, typical of a antiferromagnetic structure [30]. The hyperfine field measured by the 2 3 7 ~ P

Mossbauer effect is small (550 kOe). Specific heat data shows a very large electronic specific heat

(y = 240 mJ/mole

KZ)

and a very small magnetic entropy (0.03 R In 2), both indicating itinerant 5f electrons. Thus NpSn, seems to be an unusual case, i. e. an itinerant antiferornagnet, and the first such f electron system found.

(6)

ELECTRONIC STRUCTURE OF THE LIGHT ACTINIDES C6-311

hexagonal material shows no magnetic ordering,

although it does show some paramagnetic relaxation effects in the Mossbauer spectra and some field dependent magnetization phenomena that are diffi- cult to interpret [31]. Alloys of NpPd, with NpRh, have been investigated as well [32]. NpRh, itself does not order, although it does show behaviour which has been described by a spin-fluctuation model [33]. When Rh is added to NpPd, up to NpPd,.,,Rh

,.,,,

a decrease is observed in lattice constant, magnetic ordering temperature and paramagnetic moment. However, the hyperfiae field unexpectedly increases to 4 800 kOe, apparently in contradiction to figure l. In summary, the magnetic properties of the actinides show much variety covering many aspects of magnetic phenomena. The Mossbauer effect measurements have provided important pieces of data and should continue to provide interesting results, especially when used in conjunction with other experimental techniques.

3. Isomer shift systematics.

-

In the case of the Mossbauer resonance in ','Np and 243Am, extremely large isomer shifts are seen. For example, in ','Np the removal of a single 5f electron causes an isomer shift of roughly 25 mm/s, in comparison with the commonly observed line-widths of 2 or 3 mm/s [6, 71. Thus one has a great deal of sensitivity to changes in electronic structure. Unfortunately, chemical effects tend to be complicated since the 5f electrons participate in bonding effects in insulating compounds, and in the conduction processes in metallic systems. As m example, one would expect a rather simple situation with the available fluorides NpF, ( X = 3, 4, 5, 6).

These should be more ionic than any other compounds, and reliable relativistic calculations are available for the electron density at the nucleus in the ionic configu- rations [34]. However, comparison of the calculated densities with the observed isomer shifts show that even for this simplest of systems, simple ionic configu- rations will not be sufficient [35]. A molecular cluster calculation has been recently performed for the actinide hexafluorides [36] giving energy levels, but no calculation of the isomer shift has yet been attempted. A disadvantage of the fluorides is that they all have different crystal structures. Systematic variations in bonding may be explored more profitably through the

actinyl compounds. These consist of a strongly bound

linear molecule of the form 0-An-0 where An is a pentavalent or hexavalent actinide ion. This is a very stable structure and is a predominant feature of actinide chemistry. The uranyl radical (UO,)++ is a closed structure with all available U and 0 electrons being used to fill all available bonding orbitals. This becomes the core on which other actinyls are built. The neptunyl ion (NpO,)++ then is considered to be equivalent to a uranyl core plus one non-bonding 5f electron. This

special feature should make it possible for a study of hyperfine interactions in isomorphic uranyl and neptu-

nyl compounds to separate the effects due to bonding and non-bonding orbitals.

The An-0 bond is certainly the dominant force in actinyl compounds. However, studies of crystal che- mistry have shown that the presence of the other ligands in the compounds cause the length of the An-0 bond to vary [37]. This is in turn reflected in the hyper- fine parameters. In figure 3 we show the isomer

6 I I I I I

1.65 1.7 1.75 1.8

U-0 B O N D LENGTH I N

I S O M O R P H I C U R A N I U M C O M P O U N D S ( ~ )

FIG. 3.

-

237Np isomer shifts and quadrupole interactions in

O\lpOz)++ compounds vs. the U-0 bond length in isomorphic

(UOz)++ compounds.

shifts and quadrupole interactions for a number of neptunyl ions versus the U-0 bond length in the isomorphic U compound. The variations are clearly seen, although here too the exception to the trend is interesting. This is K3Np02F,, which is the only compound to have a strongly electronegative ligand near the actinyl radical. Thus, while one may unders- tand general features of these materials by considering the actinyl radical in isolation, a detailed discussion can not ignore the other ligands.

Here, as well, some molecular cluster calculations are becoming available. In particular, these has been a recent ESCA study of uranyl ions and a compa- rison to relativistic molecular cluster calculations [38]. At present, similar calculations for the neptunyls have met with convergence difficulties, however, these are expected to be soluble. When these produce calculated isomer shifts, a comparison with this data should be very instructive.

(7)

of a hard-sphere packing model [39]. The structure of the cubic Laves phase AB, materials is such that one is able to determine the atomic diameters dA and dB for the two components independently from the crystallo- graphic data. Ideal close-packing for hard spheres is achieved with dA/dB = 1.225. However, one finds Np and Am Laves phase compounds with this ratio varying substantially from the ideal value, indicating that significant changes in the sizes of constituent atoms are taking place [40]. These changes are usually characterized by the quantities D,

-

d, and DB

-

dB where D*, D, are the metallic atomic diameters and dA

and dB are the constituent atomic diameters within the Laves phase material as determined by lattice constant measurements. What is found is that when RA/RB

>

1.225, the actinide atoms contract [(D,

-

dA)

>

01

while the other component is roughly unchanged in size [(DB

-

dB)

-

01. Thus the B atoms set the basic size of the unit cell and the actinide ions adjust their size to fit. This change in atomic size is clearly reflected in the isomer shifts, shown in figure 4

FIG. 4.

-

Isomer shifts in Np and Am Laves phase compounds vs. the change in atomic diameter of the actinide ion.

where the isomer shifts for Np and Am Laves phase compounds are plotted vs. D,

-

d,. In addition to these simple hard sphere size arguments, one may also expect some effect on the isomer shift due to changes in the shape of the electronic distribution related to the de-localization phenomena discussed above. Again, however, we have no simple framework for including all these effects.

Such

a

hard sphere description is more difficult for other structures where the crystallography does not allow one to separate the individual components so easily. More crudely, we may simply plot the isomer shifts vs. the atomic spacing as is done in figure 5 for the NpM [41] and NpM3 1291 systems. For the NpM compound the correlation with size again seems clear. However, in the NpM3 systems one sees that NpPd, and NpRh, show anomalously large isomer shifts compared to the other compounds. This is probably related to the fact that the B sites contain d

LATTICE SPACING (A)

FIG. 5.

-

Isomer shifts in Np intermetallic compounds vs.

lattice spacing.

atoms for NpPd, and NpRh, and p atoms for the other compounds of this structure. Considerations of hybridization in an octahedral structure, again allo- wing the f orbitals to participate in the hybrid orbitals, show that a very strong bond is obtained from a sd2 f 3 combination, with no equivalently strong orbital being present when p hydridization is involved [42]. The effect of this f-d hybridization has been clearly demonstrated by relativistic band structure calculation for uranium compounds in the Cu3Au structure [43], where it is observed to play a large role in the forma- tion of the band structure. It should be noted that at present these calculations have discussed the Fermi surface properties and general band structure, but isomer shift calculations are planned for the future. In some cases, of course, a detailed interpretation of parameters is not necessary. Often one can use trends or relative changes to obtain useful information. An interesting example is the investigation by Mintz

(8)

ELECTRONIC STRUCTURE OF THE LIGHT ACTINIDES C6-3 13

lanthanides little detailed work has been done except for a few special cases [451. The resonances in 237Np and 243Am are both very similar to that of I6'Dy in that they involve

3

-+

8

nuclear levels. For the lGIDy case, static spectra (i. e. in the limit of long relaxation times) have been discussed [46], but no calculations have been performed for intermediate relaxation times (i. e. relaxation frequencies comparable with hyperfine frequencies). The major reason for this has been that computation, even for the simplest case in which the electronic angular momentum could be represented by a spin S =

+,

involved the inversion of a complex matrix with dimensionality n = 144 a large number of times for each calculated spectra, and computer time has been prohibitive. Recently we have shown [47] that such calculations can be simplified dramatically by two simplifications : (i) In a large number of problems, a procedure is available which allows a spectrum to be computed by inverting the large matrix only once, thus greatly reducing the computer time. (ii) When the hyperfine interactions have at least axial symmetry, the n = 144 matrix can be broken down into several smaller matrices, each of which can be inverted as in (i), again shortening the computational time. On the IBM 370-195, an arbitrarily complicated spectrum [48] can be calculated in 12 seconds, and a problem in axial symmetry takes 3 seconds. Thus least square fitting is feasible and the relaxation problems encountered can be handled as straightforwardly as any others.

Examples of the variety of spectra that may be obtained in 237Np are shown in figures 6 and 7. We take a hyperfine interaction

+

B[3 1;

-

I(I

+

l)] (2) where I = for both excited and ground state. We take S =

3,

appropriate to a Kramers doublet. This is known to be proper for neptunyl (NpO,)"' cases, and for NpGf in octahedral symmetry. We consider the relaxation processes to be spherical, i. e. all non- vanishing correlation functions for the lattice bath are equal, so the relaxation is characterized by a single parameter W. In figure 6 we consider cubic symmetry, with All = A, = 3 840 MHz and B = 0. For W = 0, a four line pattern is seen, exactly as occurs in l6lDy 1461 except that the ordering of line positions is different due to different nuclear moments. As relaxation increa- ses, this slowly collapses into a broad asymmetric line, and finally to a single narrow line. Figure 7 shows more complicated appearing spectra, corres- ponding to a case of axial symmetry with

All = 3 840 MHz, A, = 2 160 MHz and B = 0. Here the four lines of figure 6 are split by the non-cubic interaction into a large number of lines, and this complexity remains in the presence of relaxation broadening and collapse. One interesting point is that

L .aF

--

situations exist where a casual observationjof spectra

I I I I I I I -100 0 100

VELOCITY (mm/s)

FIG. 6.

-

Paramagnetic relaxation for 237Np with a Kramers

ion in cubic symmetry. The hyperfine parameter is taken as

A = 3 840 MHz, and various values of the relaxation frequency

W are given in the figure.

-100 0 100 -100 0 100 VELOCITY (mrnls)

FIG. 7.

-

Paramagnetic relaxation for 237Np with a Karmers

ion in axial symmetry. The hyperfine parameters are

All = 3 840 MHz, Al = 2 160 MHz and values of the relaxation

frequency W are given on the figure.

may give an incorrect isomer shift. For example note the spectra labelled 1440 MHz in figure 7 and 2 400 MHz in figure 6 , where the position of the dip

does not give the true isomer shift.

(9)

C6-3 14 B. D. DUNLAP usually present, show almost static spectra. The

strong axiality of the 0-Np-0 bond determines a highly anisotropic hyperfine interaction rather near the effective field limit. For example, in a classic neptunyl, Rb(Np02) (No3), . 6 H 2 0 , analysis of the Mossbauer data of Stone and Pillinger [49] with eq. (2)

gives A,, =

-

3 659 MHz, A, = f 64 MHz, B = 235 MHz,

q = 0. 'This values are in good agreement with those obtained by EPR [50]. In at least one case, Na(Np0,) (C2H302)3, it was found necessary to include non- axial components in the hyperfine interaction to account for the spectra 1511.

The most prominent compound which has cubic symmetry is NpF,, where the Np6 + ion is octahedrally

coordinated to the six fluorine atoms. EPR work previously gave All = A, = - 2 064 MHz [52]. The Mossbauer spectra shows only a broad, asymmetric line because the dipolar coupling is sufficiently strong to cause rather fast relaxation (as in Fig. 6). However, using the A value the data was analyzed to obtain a relaxation frequency W = 4 800 MHz [47]. More interesting results can be obtained by diluting NpF, into diamagnetic UF6, and such experiments are planned. A number of cubic compounds have been investigated where the ground state of the ion is expected to be a

r,

quartet rather than a Kramers doublet [53]. In (TMA),NpCI,, resolved spectra are seen which may be indicative of the T 8 state. In

Cs2NpC16, intermediate relaxation occurs and two braod but widely spaced lines are seen. However, the problem of calculating spectra in the presence of a T 8

state is complicated [54] and these spectra have never been analyzed. We have recently measured spectra for the compounds Cs2NpOC15 in which the near neighbour environment is an octahedron containing five Cl ions and one 0 ion. The resulting axial symmetry gives a spectrum at 4.2 K rather like the curve labelled 480 MHz of figure 7. Unfortunately, the chemistry of this compound is complicated and we have not yet obtained very reproducible results.

An interesting case of relaxation of a very different kind has been discussed by Gal et al. [55]. In this work, a source of 241Am in cubic 6-Pu metal was used with a single line absorber, and broad spectra with a strong temperature dependence were observed. These were interpreted by the authors by assuming that the 241Am decay leaves a Np5+ ion which then transforms to the N ~state at a rate sufficient to show relaxation ~ +

distortions.

All of these bits and pieces of data are very enticing, and show that a number of interesting problems involving relaxation phenomena do occur in the acti- nides. Given our current ability to calculate spectra for a wide variety of problems, we anticipate seeing more development in this area in the near future.

5. Line width problems. - Finally we consider a technical problem, but one which probably contains

interesting physics. From the known lifetime of the nuclear state in 237Np one easily calculates a natural line width for the 59.5 keV resonance of 0.07 mm/s [6,

71. However, experimentally observed line widths are always much larger than this. For the best source, consisting of

--

5 w/o Am in cubic Th metal, the linewidth is about 1.8 mm/s. For the most commonly used source of cc-Am (hexagonal-close-packed) metal, the width is about 3 mm/s. In either case, the isomer shifts and hyperfine interactions are sufficiently large that data of high accuracy can easily be obtained. None- theless, if natural linewidths could be obtained, this would be quite an extraordinary isotope. Outside of the increase in accuracy available, the increase in line intensities by more than an ordinary of magnitude would mean that the isotope could be used over a much broader temperature range than is presently possible. In spite of the interest, the problem seems to be an elusive one.

One distinct possibility for line broadening may arise from the large isomer shifts observed. All mate- rials to date have been polycrystalline and it may be that random strains throughout the materials are sufficient to cause a distribution in isomer shifts that would be seen as a uniform increase in linewidths. At present, no single crystals are available to test this possibility. However, the fact that the same line width is seen essentially independent of absorber materials indicates that the major problem is in the source. Three different source channels are possible. One involves electron capture from 2 3 7 P ~ , and has never been used due to the difficulty in preparing a sufficient quantity of activity. A second is the

P

decay of 236U, which will be discussed shortly. The third, and almost universal, source is the cc decay of 241Am. This has been highly preferred because very high specific activities can be obtained and because the source has a half-life of 458 years. The disadvantage is that the a decay is sufficiently energetic to impart a large amount of recoil energy to the daughter Np atom. As a result it is

self implanted in the host material, leaving substantial damage behind. This has frequently been suggested as the source of the extra linewidth. We have recently attempted to shed some light on this by a measurement of the temperature dependence of the emission line in Am metal [56]. While some effects are seen in the resonant intensity that may be partially attributed t o defect annealing as the temperature increases, no significant change in the linewidth is seen in the same temperature region.

An attempt has also been made to compare the linewidths obtained by preparing a source via the process 236U(n, y)237U, with the 237U proceeding to 237Np by

P

decay [57]. The feeling here was that the

P

(10)

ELECTRONIC STRUCTURE OF THE LIGHT ACTINIDES C6-3 15

this work is that the source was not annealed following the neutron irradiation. This is especially interesting in view of a recent experiment 1581 in which a single crystal of 238U0, was irradiated in a linac to produce 237U via the process 238U(y, n)237U. Spectra taken with the unannealed source showed a complicated spectrum with no clearly identifiable lines. However, annealing for one hour in H, at 800 K produced a source with a clear resonance at the Np4+ position and a linewidth of 1.5 mm/s. While this is still, of course,

much larger than the natural linewidth, it indicates that continued work with single crystal sources properly annealed may be of value. In any case, the final solu- tion for the problem of the severe line broadening in 237Np is not yet with us, and we feel this question is deserving of considerable attention.

Acknowledgment.

-

I am grateful to my many

colleagues at the Argonne National Laboratory for the collaborations represented here.

References

[l] STONE, J. A. and PILLINGER, W. L., Phys. Rev. Lett. 13 (1964) 200.

[2] The most extensive reviews appear in The Actinides : Elec- tronic Structure and Related Properties Edited by A. J. Freeman and J. B. Darby, Jr. (Academic Press, New York) 1974.

[3] For a survey, see DUNLAP, B. D. and KA~vrus, G. M., ref. [2],Vol. 1, Chap. 5 , pp. 237-302.

[4] MONARD, J. A., HURAY, P. G. and THOMSON, J. O., Phys. Rev. B 9 (1974) 2838.

[5] MEEKER, R. D., KALVIUS, G. M., DUNLAP, B. D., RUBY, S. L. and COHEN, D., Nucl. Phys. A 224 (1974) 429. [6] PILLWER, W. L. and STONE, J. A., in Mossbauer E m t

Methodology, edited by I . Gruverman (Plenum Press, New York) 1968, Vol. 4, p. 217.

[7] DUNLAP, B. D., KALVIUS, G. M., RUBY, S. L., BRODSKY, M. B. and COHEN, D., Phys. Rev. 171 (1968) 316.

[g] BoD~, D. D., WILD, J. F. and HULET, E. K., J. Znorg. Nucl. Chem. 38 (1976) 1291.

[9] KALVIWS, G. M., RUBY, S. L., DUNLAP, B. D., SHENOY, G. K., COHEN, D. and BRODSKY, M. N., Phys. Lett. B 29 (1969) 489.

[l01 FREEMAN, A. J. and KOELLING, D. D., ref. [2], Vol. 1, Chap. 2, p. 51.

[l11 The exceptions to this rule (Ce and Yb at the ends of the series and Sm and Eu in the middle) often show pro- perties that are more directly comparable with the actinides than with the other lanthanides.

[l21 HILL, H. H., in Plutonium 1970, edited by W. N. Miner (Met. Soc. AIME, New York) 1970, p. 2-9.

[l31 For example, see DONIACH, S., ref. [2], Vol. 2, p. 51 ;

BRODSKY, M. B., Phys. Rev. B 9 (1974) 1381 ;

FRADIN, F. Y., in Plutonium 1975 and Other Actinides, edited by Blank, H. and 'Lindner, R. (North Holland, Amsterdam) 1976, p. 459.

[l41 DUNLAP, B. D., BRODSKY, M. B., KALVIUS, G. M. and SHENOY, G. K., in Plutonium 1970, edited by

W. N. Miner (Met. Soc. AIME, New York) 1970, p. 331. [l51 ALDRED, A. T., DUNLAP, B. D., HARVEY, A. R., LAM, D. J.,

LANDER, G. H. and MUELLER, M. H., Phys. Rev. B 9 (1974) 3776.

1161 LAM,D. J., DUNLAP, B. D., HARVEY, A. R., MUELLER, M. H., ALDRED, A. T., NOWE, I. and LANDER, G. H., Proc. Znt. Con$ Mug., Moscow (1973) to be published.

[l71 GAL, J., KROUP, M., HADARI, Z. and NOWIK, I., Phys. Rev. (1976) to be published, and Solid State Commun. 20 (1976) 421.

[l81 DUNLAP, B. D. and LANDER, G. H., Phys. Rev. Lett. 33 (1975) 1046.

[l91 Actinide materials tend to be very hard magnetically. Hence a measurement of p by bulk techniques often gives an ambiguous answer unless high fields are utilized. The most reliable technique for obtaining p is neutron diffraction, although one must bear in mind the com- ments discussed in the next paragraph.

[20] ALDRED, A. T., DUNLAP, B. D., LAM, D. J. and NOWIK, I.,

Phys. Rev. B 10 (1974) 1011.

[21] ALDRED, A. T., DUNLAP, B. D., LAM, D. J., LANDER, G. H., MUELLER, M. H. and NOWIK, I., Phys. Rev. B 11 (1975) 530.

1221 ALDRED, A. T., LAM, D. J., HARVEY, A. R. and DUNLAP, B. D., Phys. Rev. B11 (1975) 1169.

[23] GAL, J., HADARI, Z., ATZMONY, U., BAUMINGER, E. R., NOWIK, I. and OFER, S., Phys. Rev. B 8 (1973) 1901. [24] BRODSKY, M. B. and TRAINOR, R. J., Proc. Znt. Con$ Magne-

tism, Amsterdam (1976) to be published.

[25] ALDRED, A. T., DUNLAP, B. D. and LANDER, G. M., Phys. Rev. 14 (1976) 1276.

[26] KOELLING, D. D., private communication.

[27] BRODSKY, M. B. and TRAINOR, R. J., Proc. Europhysics Conference on Itinerant Electron Magnetism, Oxford (1976) to be published.

[28] ALDRED, A. T., private communication.

[29] CAL, J., HADARI, Z., BAUMINGER, E. R., NOWIK, I. and OFER, S., Solid State Commun. 13 (1973) 647.

[30] For example, see SHENOY, G K., DUNLAP, B. D., KAL- VIWS, G. M., TOXEN, A. M. and GAMBINO, R. J., J. Appl. Phys. 41 (1970) 1317.

1311 NELLIS, W. J., HARVEY, A. R., LANDER, G. H., DUNLAP, B. D., BRODSKY, M. B., MUELLER, M. H., REDDY, T. F. and DAVIDSON, G. R., Phys. Rev. B 9 (1974) 1041. 1321 GAL, J., KROUP, M. and HADARI, Z., Solid State Commune

19 (1976) 73.

[33] NELLIS, W. J., HARVEY, A. R. and BRODSKY, M. B., AZP Conf. Proc. # 10, edited by Graham, C. D. and Rhyne, J. J. (AIP, New York) 1973, p. 1076.

1341 For a summary and references to literature, see DUNLAP, B. D., in Mossbauer Isomer Shifs, edited by Shenoy, G. K. and Wagner, F. E. (North Holland) 1976, Chap. 11. [35] DUNLAP, B. D., SHENOY, G. K., KALVIWS, G. M., COHEN, D. and MANN, J. B., in Hyperfine Interactions in Excited Nuclei, edited by Goldring, G. and Kalish, R. (Gordon and Breach, New York) 1972, Vol. 2, p. 709.

[36] KOELLING, D. D., ELLIS, D. E. and BARTLETT, R. J., J. Chem. Phys. 65 (1976) 3331.

[37] ZACHARIASEN, W. H. and PLETTINGER, H. A., Acta Crystal- logr. 12 (1959) 526.

[38] WALCH, P. F. and ELLIS, D. E., J. Chem. Phys. (1976) to be published.

[39] N ~ v m , M. V., in Electronic Structure and Alloy Chemistry of the Transition Elements, edited by Beck, P. A. (Inter- science, New York) 1963, p. 101.

[40] ALDRED, A. T., DUNLAP, B. D., LAM, D. J. and SHENOY, G. K., in Transplutonium 1975, edited by Muller, W. and Lindner, R. (North Holland, Amsterdam) 1976, p. 191 ;

LAM, D. J. and MITCHELL, A. W., J. Nucl. Mat. 44 (1972) 279.

[41] ALDRED, A. T., DUNLAP, B. D., HARVEY, A. R., LAM, D. J.,

(11)

C6-316 B. D. DUNLAP [42] EISENSTEIN, J. C., J. Chem. Phys. 25 (1956) 142.

[43] KOELLING, D. D. and FREEMAN, A. J., in Plutonium 1975 and Other Actinides, edited by Blank, H. and Lindner, R.

(North Holland, Amsterdam) 1976, p. 291.

[44] MINTZ, M. H., GAL, J., HADARI, 2. and BIXON, M., J. Phys.

& Chem. Solids 37 (1976) 1.

[45] For example, see DUNLAP, B. D., SHENOY, G. K. and ASCH, L., Phys. Rev. B 13 (1976) 18, and references therein.

[46] WICKMAN, H. H. and NOWIK, I., J. Phys. & Chem. Solids 28 (1967) 2099.

[47] SHE~OY, G. K. and DUNLAP, B. D., Phys. Rev. B 13 (1976) 1353.

[48] This is subject to the proviso that the white noise approxi- mation is appIicable. For a discussion of the implications and possible violations of this approximation, see SHENOY, G. K., DUNLAP, B. D., DATTUGUPTA, S. and ASCH. L.. Phvs. Rev. Lett.

, , .

37 (1976) 539.

.

,

[49] STONE, J. A. and PILLINGER, W. L., Symp Faraday Society 1 (1967) 77.

[50] EISENSTEIN, J. C. and PRYCE, M. H. L., Proc. R. SOC. A 229 (1955) 20.

[51] Ref. [3] p. 273.

[52] HUTCHINSON, C. A. and WEINSTOCK, B., J. Chem. Phys. 32

(1960) 56.

[53] STONE, J. A. and URRAKER, D. G., Bull. Am. Phys. Soc. 17

(1972) 337.

[54] SHENOY, G. K., DUNLAP, B. D., DATTAGUPTA and ASCH, L., J. Physique Colloq. 37 (1976) C6-85.

[55] GAL, J., HADARI, Z., BAUMINGER, E. R., NOWIK, I. and OFER, S., Solid State Commun. 15 (1974) 1805. [56] DUNLAP, B. D. and SHENOY, G. K., J. Physique ColIoq. 37

(1976) C6-55.

[57] MEEKER, R. D., DUNLAP, B. D. and COHEN, D., J. Phys.

& Chem. Solids 37 (1976) 551.

[58] OHKAWA, A., SHOJI, T. and MATSUI, K., Research Report of Laboratory of Nuclear Science, Tohoku University 7

Références

Documents relatifs

- Pour ttudier les phCnomknes d'ordre dans les alliages, on utilise habituellement des modeles simples (le modkle d'Ising le plus souvent) dans lesquels on introduit

- Simple models for the electronic structure of interstitial impurities can be determined along the same lines as previously used successfully for substitutional alloys of

F. HARTMANN-BOUTRON Laboratoire de Spectrometrie Physique U. — Nous présentons ici quelques résultats récents relatifs au traitement de la relaxa- tion par des méthodes

In order to study the first contribution, we present results for a simple single atom cluster with two specific local environments : one which simulates a disordered compact

SOFT PHONONS AND MAGNETIC ORDERING IN THE γ-PHASE TRANSITION METAL

It is concluded that the quantum spin fcc lattice with nearest neighbour anti- ferromagnetic Heisenberg and dipolar interactions ex- hibits several distinct multi-k ground

On this subject, we have pointed out that the induction of Ku is di- rectly related to the resputtering and the stress effect in the amorphous GdCo and GdFe films, respec-

Field electron emission from a regen- erative liquid metal field emitter may be considered as successive explosive electron emission due to thermal runaway like vacuum arcing / 1 /