1: \documentclass[aps,prb,preprint,groupedaddress,showpacs,amsmath,amssymb]{revtex4}
2: %\documentclass[aps,prl,preprint,superscriptaddress]{revtex4}
3: %\documentclass[aps,prl,twocolumn,groupedaddress]{revtex4}
4:
5: \usepackage{graphicx}% Include figure files
6: \usepackage{dcolumn}% Align table columns on decimal point
7: \usepackage{bm}% bold math
8:
9:
10: \begin{document}
11:
12: \title{High temperature ferromagnetism in GdFe$_2$Zn$_{20}$: large, local moments embedded in the nearly ferromagnetic Fermi liquid compound YFe$_2$Zn$_{20}$.}
13:
14: \author{S. Jia$^1$, S. L. Bud'ko$^1$, G.D. Samolyuk$^2$, P. C. Canfield$^1$}
15: \affiliation{$^1$Ames Laboratory US DOE and Department of Physics and Astronomy, Iowa State University, Ames, IA
16: 50011\\ $^2$Ames Laboratory US DOE and Department of Chemistry, Iowa State University, Ames, IA 50011}
17:
18: \date{\today}
19:
20: \begin{abstract}
21:
22: The RFe$_2$Zn$_{20}$ series manifests strongly correlated electron behavior for the non-magnetic R = Y member and
23: remarkably high temperature, ferromagnetic ordering ($T_C$ = 86 K) for the local moment bearing R = Gd member (a
24: compound that is less than 5\% atomic Gd). In contrast, the isostructural RCo$_2$Zn$_{20}$ series manifests a
25: more typical ordering temperature ($T_N$ = 5.7 K for GdCo$_2$Zn$_{20}$) and YCo$_2$Zn$_{20}$ does not show signs
26: of correlated electron behavior. Studies of R(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ (R = Gd, Y), combined with
27: bandstructure calculations for the end members, reveal that YFe$_2$Zn$_{20}$ is a nearly ferromagnetic Fermi
28: liquid and that the remarkably high $T_C$ associated with GdFe$_2$Zn$_{20}$ is the result of submerging a large
29: local moment into such a highly polarizable matrix. These results indicate that the RFe$_2$Zn$_{20}$ series, and
30: more broadly the RT$_2$Zn$_{20}$ (T = Fe, Co, Ni, Mn, Ru, Rh, Os, Ir, Pt) isostructural family of compounds, offer
31: an exceptionally promising phase space for the study of the interaction between local moment and correlated
32: electron effects near the dilute R limit.
33:
34: \end{abstract}
35:
36: \pacs{75.20.Hr, 75.30.Mb, 75.50.Cc}
37:
38: %\keywords{}
39:
40: \maketitle
41:
42: The field of condensed mater physics has been interested in the effects of electron correlations from it inception
43: \cite{mor85a,bro90a}. To this day, the properties of elemental Fe as well as Pd continue to present problems that
44: interest both experimentalists as well as theorists \cite{zel04a}. In particular materials such as Pd or Pt, that
45: are just under the Stoner limit (often referred to as nearly ferromagnetic Fermi liquids), or materials just over
46: the Stoner limit, such as ZrZn$_2$ or Sc$_3$In on the ferromagnetic side, are of particular interest
47: \cite{mor85a,bro90a,zel04a}. Of even greater interest are new examples of nearly ferromagnetic Fermi liquids that
48: can be tuned with a greater degree of ease than the pure elements: i.e. can accommodate controlled substitutions
49: on a number of unique crystallographic sites in a manner that allows for (i) a tuning of the band filling / Fermi
50: surface and (ii) the introduction of local moment bearing ions onto a unique crystallographic site. Such a
51: versatile system would open the field to a greater range of experimental studies of strongly correlated electronic
52: states as well as potentially allow for more detailed studies of quantum criticality and possibly even novel
53: superconducting states.
54:
55: In this letter we present our first results of an extensive study of the dilute, rare earth bearing, intermetallic
56: series RT$_2$Zn$_{20}$ (R = rare earth and T = transition metal) which supports a wide range of R ions for T in
57: and near the Fe, Co and Ni columns of the periodic table. In particular, in this letter we will show how
58: YFe$_2$Zn$_{20}$ is an archetypical example of a nearly ferromagnetic Fermi liquid and how, by imbedding Gd ions
59: into this highly polarizable medium, GdFe$_2$Zn$_{20}$ can have a remarkably high ferromagnetic ordering
60: temperature of 86 K, even though it contains less than 5\% atomic Gd and the Fe is not moment bearing in the
61: paramagnetic state.
62:
63: The RT$_2$Zn$_{20}$ series of compounds was discovered in 1997 by Nasch et al. \cite{nas97a} and is isostructural
64: to the cubic CeCr$_2$Al$_{20}$ structure ($Fd\bar{3}m$ space group)\cite{kri68a,thi98a}. The rare earth and
65: transition metal ions each occupy their own single, unique crystallographic sites ($8a$ and $16d$ respectively)
66: whereas the Zn ions have three unique crystallographic sites ($96g$, $48f$ and $16c$). In addition, the rare earth
67: site is one of cubic point symmetry ($\bar{4}3m$). The coordination polyhedra for R and T are comprised fully of
68: Zn, meaning that there are no R-R, T-T or R-T nearest neighbors and the shortest R-R spacing is $\sim 6$ \AA. The
69: lattice parameters for the series are $\sim 14$ \AA. Although the crystallography of this series was well
70: detailed, there have been, to date, no measurements of these compounds' physical properties. This, to some extent,
71: is not unexpected since the limited data sets available on the RT$_2$Al$_{20}$ compounds \cite{thi98a,moz98a}
72: indicated very low ordering temperatures, consistent with the very low R-concentrations.
73:
74: Single crystals of RFe$_2$Zn$_{20}$, RCo$_2$Zn$_{20}$ and R(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ (R = Gd, Y) were grown
75: from high temperature solutions \cite{can92a} rich in Zn using initial compositions of R$_2$T$_4$Zn$_{94}$. The
76: constituent elements were placed in an alumina crucible, sealed in a quartz ampule under $\sim 1/3$ atmosphere Ar
77: and heated in a box furnace to 1000$^\circ$ C and then slowly cooled to 600$^\circ$ C over up to 100 hours. The
78: resulting single crystals were large and often manifested clear [111] facets (see the inset to figure 2a below).
79: For R(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$, $x$ values are nominal values, but these are confirmed by elemental analysis
80: as well as compliance with Vegard's law, with the lattice parameter varying linearly between the $x = 0$ and $x =
81: 1$ end points. Field and temperature dependent magnetization measurements were made using Quantum Design MPMS
82: units whereas transport and specific heat measurements were made using Quantum Design PPMS units with $^3$He
83: options. The electronic band structure was calculated using the atomic sphere approximation tight binding linear
84: muffin-tin orbital (TB-LMTO-ASA) method \cite{and75a,and84a} within the local density approximation (LDA) with
85: Barth-Hedin \cite{bar72a} exchange-correlation using the experimental values of the lattice parameters. The number
86: of atoms in the reduced unit cell is 46. A mesh of 16 $\vec{k}$ points in the irreducible part of the Brillouin
87: zone (BZ) was used. The $4f$-electrons of Gd and Lu atoms were treated as a core states (polarized in the case of
88: Gd atoms).
89:
90: Figures 1 and 2 present the temperature dependent electrical resistivity, specific heat and low field
91: magnetization data, as well as anisotropic $M(H)$ data, for GdFe$_2$Zn$_{20}$ and GdCo$_2$Zn$_{20}$. There are
92: two conspicuous differences between the physical properties of these compounds: (i) GdFe$_2$Zn$_{20}$ orders
93: ferromagnetically whereas GdCo$_2$Zn$_{20}$ orders antiferromagnetically, and (ii) GeFe$_2$Zn$_{20}$ orders at a
94: remarkably high temperature of $T_C = 86$ K whereas GdCo$_2$Zn$_{20}$ orders at the more representative $T_N =
95: 5.7$ K. From figure 2a the high temperature Curie constant can be determined, giving effective moments ($8.05
96: \mu_B$ and $8.15 \mu_B$ for T=Fe and T=Co respectively) consistent with the effective moment of Gd$^{3+}$,
97: indicating that, in the paramagnetic state, there is little or no contribution from the transition metal. The
98: saturated moment deduced from the data in figure 2b is close to that associated with Gd$^{3+}$; slightly lower for
99: GdFe$_2$Zn$_{20}$ and slightly higher for GdCo$_2$Zn$_{20}$. The magnetic entropy associated with each phase
100: transitions is approximately $R \ln 8$, with more uncertainty associated with the subtraction of the
101: YT$_2$Zn$_{20}$ data for T = Fe than for T = Co due to much higher ordering temperature. The remarkably high
102: ordering temperature found for GdFe$_2$Zn$_{20}$ is not unique to the R = Gd member of the RFe$_2$Zn$_{20}$
103: series. For R = Gd -– Tm transitions to ferromagnetically ordered states occur at temperatures that roughly scale
104: with the de Gennes parameter. \cite{jai06a}
105:
106: In order to better understand this conspicuous difference in ordering temperatures, bandstructural calculations
107: were carried out. Figure 3 presents the density of states as a function of energy for both LuFe$_2$Zn$_{20}$ and
108: LuCo$_2$Zn$_{20}$. The upper curve in each panel is the total density of states whereas the lower curve is the
109: density of states associate with the transition metal ion. It should be noted that the difference between
110: LuFe$_2$Zn$_{20}$ and LuCo$_2$Zn$_{20}$ density of states can be rationalized in terms of the rigid band
111: approximation, with the Fermi level for LuCo$_2$Zn$_{20}$ being $\sim 0.3$ eV higher than that for
112: LuFe$_2$Zn$_{20}$, associated with the two extra electrons per formula unit. Calculations done on YFe$_2$Zn$_{20}$
113: and GdFe$_2$Zn$_{20}$ as well as on YCo$_2$Zn$_{20}$ and GdCo$_2$Zn$_{20}$ lead to similar density of states
114: curves \cite{jai06a} and further analysis of the GdFe$_2$Zn$_{20}$ and GdCo$_2$Zn$_{20}$ banstructural results
115: leads to the prediction that for GdFe$_2$Zn$_{20}$ the ground state will be ferromagnetic with a total saturated
116: moment of approximately 6.8 $\mu_B$ (with a small induced moment on the Fe opposing the Gd moment) and for
117: GdCo$_2$Zn$_{20}$ the ground state will be antiferromagnetic with a saturated moment of 7.15 $\mu_B$ (and
118: practically no induced moment on Co). These results are consistent with the saturated values of the magnetization
119: seen in Fig. 2b.
120:
121: These calculations indicate that the RFe$_2$Zn$_{20}$ compounds should manifest a higher $N(E_F)$ than the
122: RCo$_2$Zn$_{20}$ analogues and bring up the question of whether or not this is the primary reason for the
123: remarkably high $T_C$ found for GdFe$_2$Zn$_{20}$. In addition they bring up the question of how correlated the
124: electronic state is in the nominally non-magnetic Lu- or Y- based analogues. In order to address these questions
125: two substitutional series were grown: Y(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ and Gd(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$.
126: Figure 4 presents thermodynamic data taken on the Y(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ series. For $x = 0$ the low
127: temperature, linear component of the specific heat ($\gamma$) is relatively small (19 mJ/mol K$^2$) and the
128: susceptibility is weakly paramagnetic and essentially temperature independent. As $x$ is increased there is a
129: monotonic (but clearly super-linear) increase in the samples' paramagnetism as well as, for larger $x$-values, an
130: increase in the low temperature $\gamma$ values. For YFe$_2$Zn$_{20}$ ($x = 1$) the value of gamma has increased
131: to over 250\% of that for YCo$_2$Zn$_{20}$ and the susceptibility has become both large and temperature dependent.
132: Figure 5 presents data on the analogous Gd(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ series. As $x$ is increased the ground
133: state rapidly becomes ferromagnetic and the transition temperature increases monotonically (but again in a
134: super-linear fashion) and the high field, saturated, magnetization decreases weakly, in a monotonic fashion.
135:
136: Taken together, figures 4 and 5 demonstrate a clear correlation between $x$, the linear component of the
137: electronic specific heat, the enhanced magnetic susceptibility of the Y-based series and the Curie temperature and
138: the saturated magnetization of the Gd-based series. This correlation can be more clearly seen if the relation
139: between the linear component of the specific heat and the low temperature susceptibility of the Y-based series is
140: placed in the context of a nearly ferromagnetic Fermi liquid: i.e. if the Stoner enhancement parameter, $Z$, for
141: each member of the series can be determined.\cite{zim72a} For such systems the static susceptibility (corrected
142: for the core diamagnetism \cite{cd}) is $\chi = \chi_0/(1-Z)$, where $\chi_0 = \mu_B N(E_F)$ is the Pauli
143: paramagnetism. Given that the linear component of the specific heat is given by $\gamma_0 = (\pi k_B)^2 N(E_F)/3$,
144: if both the low temperature specific heat and magnetic susceptibility can be measured, then the parameter $Z$ can
145: be deduced, ($Z = 1 - (3\mu_0^2/\pi k_B^2)(\gamma_0/\chi_0)$). The canonical example of such a system is elemental
146: Pd for which, using data from ref. \onlinecite{cho68a}, $Z = 0.83$. For YFe$_2$Zn$_{20}$, $Z = 0.89$, a value
147: that places it even closer to the Stoner limit that Pd. It should be noted that the temperature dependent
148: susceptibility of YFe$_2$Zn$_{20}$ is remarkably similar to that of Pd as well (see ref. \onlinecite{zel04a} and
149: references therein). The $x$-dependence of the experimentally determined values of $\gamma$ and $\chi(T = 0)$, as
150: well as the inferred value of $Z$, for the Y(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ series are plotted in Fig. 6a. By
151: choosing $x$, Y(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ can be tuned from being exceptionally close to the Stoner limit to
152: well removed from it. Corrections to these inferred $Z$ values coming from the difference between the measured
153: electronic specific heat coefficient, $\gamma$, and Sommerfeld coefficient, $\gamma_0$, where $\gamma = \gamma_0
154: (1+\lambda)$ only serve to slightly increase $Z$ since $\lambda$, the electron mass enhancement parameter, is
155: positive definitive. We can estimate $\lambda = 0.85$ and 0.22, for $x = 1$ and $x=0$ respectively, and this
156: shifts $Z$ to 0.94 for YFe$_2$Zn$_{20}$ and to 0.50 for YCo$_2$Zn$_{20}$.
157:
158: When the non-magnetic Y ion is replaced by the large, Heisenberg moment associated with the $S = 7/2$ Gd$^{3+}$
159: ion, as $x$ is varied from zero to one in the Gd(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ series, the Gd local moments will
160: be in an increasingly polarizable matrix, one that is becoming a nearly ferromagnetic Fermi liquid. This results
161: in an increasingly strong coupling between the Gd local moments as $x$ is increased. Figure 6b plots the
162: $x$-dependence of $T_C$ and $\mu_{sat}$ for the Gd(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$. The value of $T_C$ increases
163: in a monotonic but highly non-linear fashion in a manner reminiscent to the behavior associated with the
164: increasingly polarizability of Y(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ seen in figure 6a. The $\mu_{sat}$ value for the
165: field applied along the [111] direction varies systematically from the slightly enhanced 7.3 $\mu_B$ value found
166: for GdCo$_2$Zn$_{20}$ to the slightly deficient value of 6.5 $\mu_B$ found for GdFe$_2$Zn$_{20}$.
167:
168: In addition to $x$-dependence, this conceptually simple and compelling framework can also be used to understand
169: the rather curious temperature dependence of the $1/\chi$ vs. $T$ data for GdFe$_2$Zn$_{20}$ seen in Fig. 2a. As
170: $T$ is decreased the electronic background that the Gd$^{3+}$ ion is immersed in becomes increasingly polarizable
171: (as shown in Fig. 4a for $x = 1$), leading to a temperature dependent coupling that in turn leads to the
172: non-linearity of the $1/\chi$ vs. $T$ data. If a constant effective moment for the Gd$^{3+}$ is assumed, then a
173: temperature dependent, paramagnetic $\Theta$ can be extracted from the data in Fig. 2a: $\chi(T) = \chi_0 + C/(T
174: + \Theta)$. This temperature dependent $\Theta$, shown in Fig. 4a, tracks the electronic susceptibility of the
175: YFe$_2$Zn$_{20}$ remarkably well, both increasing by $\sim 1.7$ upon cooling from 300 K to 100 K.
176:
177: One consequence of placing Gd ions into a matrix so close to the Stoner limit is an enhanced sensitivity to small
178: sample-to-sample variations. This is most clearly illustrated by the data for the
179: Gd(Fe$_{0.88}$Co$_{0.12}$)$_2$Zn$_{20}$ samples shown in Figs. 5a and 6b. Although the samples have the same
180: nominal composition there is a clear difference in their transition temperatures (Fig. 5a). This difference
181: though is not too significant given the large $dT_C/dx$ slope seen in Fig. 6b. On the other hand, measurements on
182: four separate samples of Gd(Fe$_{0.25}$Co$_{0.75}$)$_2$Zn$_{20}$ did not show any significant variations in $T_C$.
183: Such sensitivity of correlated electron states to small disorder is not uncommon, giving rise to significant
184: variation in measured $T_C$ for samples of Sc$_3$In and ZrZn$_2$ [\onlinecite{mor85a,flo06a}] as well as dramatic
185: changes in the transport properties of heavy fermions such as YbNi$_2$B$_2$C. \cite{avi02a}
186:
187: In summary, the R(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ series offers a unique opportunity to systematically study the
188: evolution of a nearly ferromagnetic Fermi liquid (for R = Y, Lu) and also offers the possibility of studying the
189: effects that such a polarizable background has on the ordering of large local moments which are located on an
190: existing, unique crystallographic site. The broader RT$_2$Zn$_{20}$ family of compounds offers an even larger
191: phase space for the study of correlated electron physics (for T = Fe, Ru, and Os as well as for R = Yb and Ce)
192: \cite{jai06a,tor06a} and for the study of local moment physics, all in the limit of a dilute, rare earth bearing,
193: intermetallic series. The study of these compounds promises to be a fruitful new phase space for several years to
194: come.
195:
196:
197: \begin{acknowledgments}
198: We are indebted to the following students and magneticians: K. Dennis, N. Ni, J. Friedrich, S.A. Law, H. Ko, E.D.
199: Mun, A. Safa-Sefat for help in samples' growth and characterization and to J. Schmalian and B.N. Harmon for useful
200: discussions. Ames Laboratory is operated for the U.S. Department of Energy by Iowa State University under Contract
201: No. W-7405-Eng.-82. This work was supported by the Director for Energy Research, Office of Basic Energy Sciences.
202: \end{acknowledgments}
203:
204:
205: \begin{references}
206:
207: \bibitem{mor85a} T. Moryia, {\it Spin fluctuations in itinerant electron magnetism} (Springer, Berlin, 1985).
208:
209: \bibitem{bro90a} P.E. Brommer, and J.J.M. Franse, in: {\it Ferromagnetic Materials}, Vol. 5, edited by K.H.J. Buschow and E.P. Wohlfarth (Elsevier, Amsterdam, 1990) p.224.
210:
211: \bibitem{zel04a} B. Zellermann, A. Paintner, and J. Voitl$\ddot{a}$nder, {\it J. Phys.: Cond. Mat.} {\bf 16} (2004) 919.
212:
213: \bibitem{nas97a} T. Nasch, W. Jeitschko, and U.C. Rodewald, {\it Zeitschrift f$\ddot{u}$r Naturforschung, B: Chem. Sciences} {\bf 52}
214: (1997) 1023 .
215:
216: \bibitem{kri68a} P.I. Kripyakevich, and O.S. Zarechnyuk, {\it Dopov. Akad. Nauk Ukr. RSR, Ser. A} {\bf 30} (1968) 364.
217:
218: \bibitem{thi98a} V.M.T. Thiede, W. Jeitschko, S. Niemann, and T. Ebel, {\it J. Alloys Compd.} {\bf 267} (1998) 23.
219:
220: \bibitem{moz98a} O. Moze, L.D. Tung, J.J.M. Franse and K.H.J. Buschow, {\it J. Alloys Compd.} {\bf 268} (1998) 39.
221:
222: \bibitem{can92a} P.C. Canfield and Z. Fisk, {\it Phil. Mag. B} {\bf 65} (1992) 1117.
223:
224: \bibitem{and75a} O.K. Andersen, {\it Phys. Rev. B} {\bf 12} (1975) 3060.
225:
226: \bibitem{and84a} O.K. Andersen, O. Jepsen, {\it Phys. Rev. Lett.} {\bf 53} (1984) 2571.
227:
228: \bibitem{bar72a} U. von Barth, L. Hedin, {\it J. Phys. C} {\bf 5} (1972) 1629.
229:
230: \bibitem{jai06a} S. Jia, {\it et al.}, unpublished.
231:
232: \bibitem{zim72a} J.E. Ziman, {\it Principles of the theory of solids}, 2nd edition (Cambridge University Press, Cambridge, 1972).
233:
234: \bibitem{cd} Using Table 5.7 (p. 306) in L.N. Mulay, E.A. Boudreaux {\it Theory and applications of molecular diamagnetism} (John Wiley \& Sons, NY 1976) we can estimate the core diamagnetism of Y(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ to be $-2.3 \times 10^{-4}$ emu/mol.
235:
236: \bibitem{cho68a} G. Chouteau, R. Fourneaux, K. Gobrecht, and R. Tournier, {\it Phys. Rev. Lett.} {\bf 20} (1968)
237: 193.
238:
239: \bibitem{flo06a} J. Flouquet, {\it et al.}, private communication.
240:
241: \bibitem{avi02a} M.A. Avila, S.L. Bud'ko, and P.C. Canfield {\it Phys. Rev. B} {\bf 66} (2002) 132504;
242: M.A. Avila, Y.Q. Wu, C.L. Condron, S.L. Bud'ko, M. Kramer, G.J. Miller, and P.C. Canfield {\it Phys. Rev. B} {\bf
243: 69} (2004) 205107; S.L. Bud'ko and P.C. Canfield {\it Phys. Rev. B} {\bf 71} (2005) 024409.
244:
245: \bibitem{tor06a} M. S. Torikachvili, S. Jia, S. T. Hannahs, R. C. Black, W. K. Neils, Dinesh Martien, S. L. Bud'ko, and P. C. Canfield, cond-mat/0608422.
246:
247: \end{references}
248:
249: \clearpage
250:
251: \begin{figure}
252: \begin{center}
253: \includegraphics[angle=0,width=120mm]{Pig1a.eps}
254: \includegraphics[angle=0,width=120mm]{Pig1b.eps}
255: \end{center}
256: \caption{Temperature dependent specific heat, resistivity, and low field ($H = 1000$ Oe) magnetization divided by
257: field for (a) GdFe$_2$Zn$_{20}$ and (b) GdCo$_2$Zn$_{20}$. For GeFe$_2$Zn$_{20}$ $\rho(300 K) = 73~\mu\Omega$ cm
258: and for GdCo$_2$Zn$_{20}$ $\rho(300K) = 60~\mu\Omega$ cm.}\label{F1}
259: \end{figure}
260:
261: \clearpage
262:
263: \begin{figure}
264: \begin{center}
265: \includegraphics[angle=0,width=100mm]{Pi2a.eps}
266: \includegraphics[angle=0,width=100mm]{Pig2b.eps}
267: \end{center}
268: \caption{(a) $H/M$ as a function of $T$ for GdFe$_2$Zn$_{20}$ and GdCo$_2$Zn$_{20}$ (dotted lines show region used
269: for the high temperature Curie-Weiss fit). (b) Anisotropic $M(H)$ ($T = 1.85$ K) for GdFe$_2$Zn$_{20}$ and
270: GdCo$_2$Zn$_{20}$. For each sample measurements for $H\|[100]$, $H\|[110]$, and $H\|[111]$ are shown. Inset to
271: figure 2a: a single crystal of YFe$_2$Zn$_{20}$ next to a mm scale.}\label{F2}
272: \end{figure}
273:
274: \clearpage
275:
276: \begin{figure}
277: \begin{center}
278: \includegraphics[angle=270,width=120mm]{Pig3A.ps}
279: \end{center}
280: \caption{Density of states as a function of energy for LuFe$_2$Zn$_{20}$ and LuCo$_2$Zn$_{20}$. The upper curve
281: is total density whereas the lower curve is partial density of states associated with Fe or Co.}\label{F3}
282: \end{figure}
283:
284: \clearpage
285:
286: \begin{figure}
287: \begin{center}
288: \includegraphics[angle=0,width=120mm]{Pig4a.eps}
289: \includegraphics[angle=0,width=120mm]{Pig4b.eps}
290: \end{center}
291: \caption{Temperature dependent magnetic susceptibility (a) and low temperature $C/T$ as a function of $T^2$ (b)
292: for Y(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ series. Inset to figure 4a: temperature dependence of paramagnetic $\Theta$
293: for GdFe$_2$Zn$_{20}$ (see text).}\label{F4}
294: \end{figure}
295:
296: \clearpage
297:
298: \begin{figure}
299: \begin{center}
300: \includegraphics[angle=0,width=100mm]{Pig5a.eps}
301: \includegraphics[angle=0,width=100mm]{Pig5b.eps}
302: \end{center}
303: \caption{Low field magnetization divided by field as a function of temperature for $x$ = 1.00, 0.88, 0.75, 0.50,
304: 0.25, and 0.00 - from right to left (a) and low temperature ($T = 1.85$ K) magnetization as a function of applied
305: field (b) for Gd(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ series. Note that in (a) data from two samples of $x = 0.88$ are
306: shown. }\label{F5}
307: \end{figure}
308:
309: \clearpage
310:
311: \begin{figure}
312: \begin{center}
313: \includegraphics[angle=0,width=100mm]{Pig6aB.eps}
314: \includegraphics[angle=0,width=100mm]{Pig6b.eps}
315: \end{center}
316: \caption{Plots of linear coefficient of the specific heat, $\gamma$, magnetic susceptibility for $T \to 0$
317: corrected for core diamagnetism \cite{cd}, $\chi_0$, and Stoner enhancement factor, $Z$, for
318: Y(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ series (a) and $T_C$ for Gd(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ series (b). Inset to
319: (b) shows $M_{sat}$ as a function of $x$ for Gd(Fe$_x$Co$_{1-x}$)$_2$Zn$_{20}$ series.}\label{F6}
320: \end{figure}
321:
322:
323: \end{document}
324: