1: The numerical renormalization (NRG) was developed by Wilson
2: as a numerical approach to the single-impurity Kondo and Anderson
3: problems, which, in conjunction with various analytic methods,
4: provides an accurate, complete solution of the problems
5: \cite{wilson_rg}.
6: The difficulty in these problems lies in the widely varying energy
7: scales that need to be accurately described.
8: For example, Kondo screening occurs at an exponentially small energy
9: scale set by the Kondo temperature.
10: Perturbation theory, while providing a good description at higher
11: temperatures, breaks down at the Kondo temperature.
12:
13: Here we will be concerned primarily with illustrating the general
14: concepts and features of the method, exploring the conditions under
15: which it works or does not work, and relating it to other
16: methods; in particular, to exact diagonalization and to the DMRG.
17: More complete treatments of the NRG include
18: Refs.\ \cite{wilson_rg,hewson_book,costi_dmrg_book}, on which the material
19: presented here is based.
20:
21: \subsection{Anderson and Kondo problems}
22: The single-impurity Anderson model (SIAM) was introduced by Anderson
23: in 1961 to
24: describe a spherically symmetric strongly correlated impurity in an
25: uncorrelated non-magnetic metal \cite{anderson}.
26: In lattice form, its Hamiltonian can be written
27: \begin{equation}
28: H^{AI} \; = \; \varepsilon_d \, \sum_\sigma n^d_\sigma
29: \; + \; U n^d_{\uparrow} n^d_{\downarrow}
30: \; + \; \sum_{{\bf k},\sigma} \left ( V_{{\bf k},d} \,
31: c^\dagger_{{\bf k}, \sigma} d^{\phantom{\dagger}}_{{\bf k}, \sigma} +
32: {\rm H.c.}
33: \right )
34: \; + \; \sum_{{\bf k},\sigma} \varepsilon_{{\bf k}}\,
35: c^\dagger_{{\bf k},\sigma} c^{\phantom{\dagger}}_{{\bf k},\sigma}
36: \label{eqn:SIAM_chain}
37: \end{equation}
38: where the non-degenerate impurity has energy $\varepsilon_d$ and
39: $n^d_\sigma = d^\dagger_{\sigma} d^{\phantom{\dagger}}_\sigma$
40: is its number operator and
41: $d^\dagger_{\sigma}$ the corresponding creation operator.
42: The Coulomb interaction $U$ is local and takes place only on the
43: impurity, but the hybridization $V_{{\bf k},d}$ in general allows
44: scattering from all momentum states.
45: The hybridization function
46: \begin{equation}
47: \Delta(\omega) = \pi \sum_{\bf k}|V_{{\bf k},d}|^2
48: \delta(\omega - \varepsilon_{\bf k} ) \; ,
49: \end{equation}
50: along with the density of states of the conduction band,
51: $\rho(\omega) = \sum_{\bf k}\delta(\omega - \varepsilon_{\bf k} )$,
52: determine the behavior of the system.
53: Here we follow the usual treatment and consider only the orbitally
54: symmetric case, i.e., $V_{{\bf k},d} = V_{k d}$ and
55: $\varepsilon_{\bf k} = \varepsilon_k$.
56: This corresponds to taking a completely isotropic electron gas, which
57: can be realized only approximately in a real solid.
58: We then need to consider only the s-wave states of the conduction
59: electrons, so that $c_{\bf k, \sigma}$ can be replaced by
60: $c_{k, l=m=0, \sigma} \equiv c_{k, \sigma}$.
61:
62:
63: A closely related model is the Kondo or s-d exchange model,
64: \begin{equation}
65: H^{K} \, = \, J_K \, {\bf S}_d \cdot {\bf s}_0
66: \, + \, \sum_{{\bf k},\sigma} \varepsilon_{{\bf k}}\,
67: c^\dagger_{{\bf k},\sigma} c^{\phantom{\dagger}}_{{\bf k},\sigma}
68: \label{eqn:kondo}
69: \end{equation}
70: with
71: $
72: {\bf s}_0 = f^{\dagger}_{0,\sigma} {\vec{\sigma}}_{\sigma\mu}
73: f^{\phantom{\dagger}}_{0,\mu}
74: $ and localized Wannier state generated by
75: $f^{\phantom{\dagger}}_{0,\sigma} =
76: \sum_k c^{\phantom{\dagger}}_{k,\sigma}$,
77: in which the electronic impurity is replaced by a localized spin-1/2
78: described by the spin operator ${\bf S}_d$.
79: The Kondo model can be derived as a strong coupling ($U \gg V$)
80: approximation to the SIAM via the Schrieffer-Wolff
81: transformation in the symmetric case ($\varepsilon_d = -U/2$)
82: \cite{schrieffer-wolff}, leading to the correspondence
83: $J_K = 8 V^2/U$.
84:
85: In order to bring the SIAM and the Kondo model into a form amenable to
86: numerical treatment, the models are mapped onto a linear chain model
87: using a Lanczos tridiagonalization procedure.
88: For the SIAM with orbital symmetry, the procedure is carried out as
89: follows.
90: The object is to transform the hybridization into a local term.
91: This can be done by defining
92: the localized Wannier state $|0,\sigma \rangle \equiv
93: f^\dagger_{0,\sigma} |\psi_{\rm vac} \rangle$ ($|\psi_{\rm vac}
94: \rangle$ designates the vacuum) with
95: \begin{equation}
96: f_{0,\sigma} = \frac{1}{V}\sum_k V_{k d} c_{k,\sigma}
97: \end{equation}
98: and the normalization $V \equiv \left (\sum_k |V_{k d}|^2 \right )^{1/2}$.
99: One then transforms the kinetic energy of the conduction band,
100: $H_c = \sum_k \varepsilon_k
101: c^\dagger_{ k,\sigma} c^{\phantom{\dagger}}_{k,\sigma}$ to the new set
102: of operators, taking it from diagonal %(!)
103: to tridiagonal form, by constructing a sequence of orthogonal states
104: generated by applying $H_c$:
105: \begin{eqnarray}
106: |1 \rangle & = & \frac{1}{\lambda_0}
107: \left [ H_c | 0 \rangle - |0 \rangle \langle 0 | H_c | 0 \rangle
108: \right ]\nonumber\\
109: %\ldots & & \\
110: |n + 1 \rangle & = & \frac{1}{\lambda_n}
111: \left [ H_c | n \rangle - |n \rangle \langle n | H_c | n \rangle
112: - |n -1 \rangle \langle n-1 | H_c | n \rangle \rangle \right ]
113: \end{eqnarray}
114: (where we have dropped the spin index $\sigma$ for compactness).
115: Note that this is a rather unusual analytic application of what is
116: essentially the Lanzcos procedure of Section \ref{lanczos_method}.
117: The result is that the SIAM is transformed to the form of a
118: semi-infinite tight-binding chain
119: \begin{eqnarray}
120: \tilde{H}^{AI} \; = \; \varepsilon_d \, n^d_\ell
121: \;& + &\; U n^d_{\ell, \downarrow} n^d_{\ell, \downarrow}
122: \; + \; V \sum_{{\bf k},\sigma}\, \left (
123: f^\dagger_{0, \sigma} d^{\phantom{\dagger}}_{\sigma} +
124: {\rm H.c.} \right ) \nonumber \\
125: \; & + & \; \sum_{n=0,\sigma}^\infty \left [ \epsilon_{n}\,
126: f^\dagger_{n,\sigma} f^{\phantom{\dagger}}_{n,\sigma}
127: \, + \, \lambda_n \left (
128: f^\dagger_{n, \sigma} f^{\phantom{\dagger}}_{n+1,\sigma} +
129: {\rm H.c.} \right )
130: \right ]
131: \; ,
132: \end{eqnarray}
133: where, in second quantized notation, the operators
134: $f^\dagger_{n,\sigma}$ create an electron in state $|n \rangle$.
135: The Lanczos coefficients, $\lambda_n = \langle n+1 | H_c| n\rangle$
136: and $\epsilon_n = \langle n | H_c| n \rangle$, depend
137: on the dispersion $\varepsilon_k$ and the hybridization function
138: $\Delta(\omega)$ of the original SIAM, and, in general, can be
139: determined, in the worst case numerically, during the Lanczos
140: procedure.
141: In general, they do not necessarily fall off with $n$, a property which is
142: crucial for the convergence of the NRG procedure, as we will see
143: below.
144: An analogous procedure can be carried out for the Kondo Hamiltonian
145: (\ref{eqn:kondo}), starting with the local Wannier state
146: $|0, \sigma\rangle$ generated by
147: $f^{\phantom{\dagger}}_{0,\sigma} =
148: \sum_k c^{\phantom{\dagger}}_{k,\sigma}$.
149:
150: In order to ensure that the tridiagonal chain formulations of the
151: SIAM or the Kondo model lead to a convergent NRG
152: procedure, additional approximations must be made.
153: In particular, a logarithmic discretization of the conduction band
154: leads to coefficients that fall off exponentially with $n$.
155: Here, we will discuss the Kondo model for concreteness, but extension
156: to the SIAM is straightforward.
157: In the first step, we take the density of states of the conduction
158: band to be a constant, $D$.
159: As long as there are no divergences at the Fermi level, the low-energy
160: physics should be dominated by the constant part, as can be justified by
161: expanding the dispersion $\varepsilon_k$ in a power series about the
162: Fermi vector $k_F$ \cite{wilson_rg}.
163: (Generalizations can be made for other densities of states.)
164: The next is a logarithmic discretization of the conduction
165: band $[-D < \varepsilon_k < D ]$ in energy, i.e., a division into
166: intervals $D^+ = [\Lambda^{-(n+1)},\Lambda^{-n}]$ and
167: $D^- = [-\Lambda^{-n},-\Lambda^{-(n+1)}]$ for the positive and negative
168: parts, respectively (see Fig.\ \ref{fig:log_discret}).
169: For each interval, we can expand the electron creation operators in a
170: Fourier series, defining
171: \begin{equation}
172: a^\dagger_{n,p,\sigma} \equiv \sum_{[k]_n} e^{i \omega_n p k }
173: c^\dagger_{k,\sigma} \; ,
174: \end{equation}
175: where $[k]_n$ is the set of momentum points in the interval
176: $\Lambda^{-(n+1)} < k < \Lambda^{-n}$, $p$ is an integer, and
177: $\omega_n = 2 \pi \Lambda^{n+1}/(\Lambda-1)$.
178: For an interval $\Lambda^{-(n+1)} < k < \Lambda^{-n}$ in the
179: negative range, operators $b^\dagger_{n,p,\sigma}$ can be defined
180: analogously,
181: It can be shown that the operators $a^\dagger_{n,p,\sigma}$,
182: $b^\dagger_{n,p,\sigma}$, and their hermitian conjugates obey the usual
183: anticommutation rules for fermions.
184: For $H^K$, Eq.\ (\ref{eqn:kondo}), the localized state can be
185: expressed as
186: \begin{equation}
187: f^\dagger_{0,\sigma} =
188: \left ( 1 - \Lambda^{-1} \right )^{1/2} \;
189: \sum_n \Lambda^{-n/2} \;
190: \left ( a^\dagger_{0,p,\sigma} + b^\dagger_{0,p,\sigma} \right )
191: \end{equation}
192: At this point, the approximation is made to neglect all higher terms
193: in the Fourier series, keeping only one electron per logarithmic
194: energy interval, i.e.,
195: $a^\dagger_{n,0,\sigma}$ and $b^\dagger_{n,0,\sigma}$.
196: Since the impurity only couples directly to the localized state,
197: neglecting these states amounts only to neglecting off-diagonal matrix
198: elements, which can be shown to be proportional to
199: $(1-\Lambda^{-1})$.
200: Therefore, the approximation becomes valid in the limit $\Lambda\to 1$;
201: a more detailed analysis of the errors can be found in
202: Ref.\ \cite{wilson_rg}.
203:
204: \begin{figure}
205: \includegraphics[width=0.6\textwidth]{NRG_figs/log_discret.eps}
206: \caption{logarithmic discretization of the conduction band.}
207: \label{fig:log_discret}
208: \end{figure}
209:
210: After applying the tridiagonalization procedure outlined above to the
211: logarithmically discretized version of the Hamiltonian $H^K$, we
212: arrive at the effective chain Hamiltonian for the Kondo
213: model
214: \begin{equation}
215: \tilde{H}^K = \tilde{D} \, \sum_{n=0, \sigma}^\infty \,
216: \Lambda^{-n/2} \, \left (
217: f^\dagger_{n, \sigma} f^{\phantom{\dagger}}_{n+1,\sigma} +
218: {\rm H.c.} \right )
219: \, + \,
220: 2 J_K \, \sum_{\sigma, \mu} \, f^\dagger_{0, \sigma} {\vec{\sigma}}_{\sigma\mu}
221: f^{\phantom{\dagger}}_{0,\mu} \;
222: \label{eqn:kondo_chain}
223: \end{equation}
224: where $\tilde{D} = D (1 + \Lambda^{-1})/2$.
225: The tight-binding part has the same form as that in
226: Eq.\ (\ref{eqn:SIAM_chain}) with
227: $\epsilon_n = 0$ and
228: $\lambda_n \approx \frac{1}{2} (1+\Lambda^{-1})\Lambda^{-n/2}$ for
229: large $n$ \cite{wilson_rg}.
230:
231: This form of the tight-binding Hamiltonian for the Kondo model and an
232: analogous one for the SIAM are treatable numerically with the NRG.
233: Crucial for the convergence is that the tight-binding coupling decays
234: exponentially with the position on the lattice.
235: Physically, this discretization was carefully thought out by Wilson to
236: reflect the exponentially small energy scales evident in the behavior
237: of perturbation theory for the Kondo problem \cite{wilson_rg,hewson_book}.
238: Note that it is crucial to adjust the discretization parameter
239: appropriately.
240: If $\Lambda$ is too close to unity, the NRG will not converge
241: sufficiently quickly, if $\Lambda$ is chosen to be too large, the error
242: from the logarithmic discretization becomes too large.
243: In practice, one chooses a value around 2, but the extrapolation
244: $\Lambda\to1$ should, in principle, be carried out.
245:
246: \subsection{Numerical RG for the Kondo Problem}
247:
248: \subsubsection{Renormalization Group Transformation}
249:
250: We will now outline the NRG procedure as applied to Hamiltonian
251: (\ref{eqn:kondo_chain}).
252: The goal is to investigate the behavior of the system at a given
253: energy scale by treating a finite system of length $L$ with
254: Hamiltonian
255: \begin{equation}
256: \tilde{H}_L = \tilde{D} \, \sum_{n=0, \sigma}^{L-1} \,
257: \Lambda^{-n/2} \, \left (
258: f^\dagger_{n, \sigma} f^{\phantom{\dagger}}_{n+1,\sigma} +
259: {\rm H.c.} \right )
260: \, + \,
261: 2 J \, \sum_{\sigma, \mu} \, f^\dagger_{0, \sigma} {\vec{\sigma}}_{\sigma\mu}
262: f^{\phantom{\dagger}}_{0,\mu}
263: \end{equation}
264: (dropping the superscript/subscript ``K'' for compactness).
265: Due to the exponentially decaying couplings, a particular system size will
266: then describe the energy scale set by
267: $D_L \equiv \tilde{D}/\Lambda^{(L-1)/2}$.
268:
269: The idea of Wilson was to examine the behavior (i.e., the renormalization
270: group flow) of the lower part of the appropriately rescaled eigenvalue
271: spectrum of the sequence of Hamiltonians
272: $\tilde{H}_L$, $\tilde{H}_{L+1}$, $\ldots$ numerically.
273: It is convenient to rescale the Hamiltonians directly, defining
274: $H_L \equiv \tilde{H}_L/D_L$.
275: One can relate $H_L$ to $H_{L+1}$ through the recursion relation
276: \begin{equation}
277: H_{L+1} = \Lambda^{1/2} \, H_L + \sum_{\sigma}
278: \left ( f^\dagger_{L, \sigma} f^{\phantom{\dagger}}_{L+1,\sigma} +
279: {\rm H.c.} \right )
280: \; \equiv {\cal R}[H_L] \; .
281: \label{eqn:NRG_recursion}
282: \end{equation}
283: This defines the RG transformation.
284: In principle, this transformation is exact (up to the discretization
285: error associated with $\Lambda$).
286: However, if one were to treat each subsequent $H_L$ by numerical
287: diagonalization, the memory and work needed would increase
288: exponentially because the number of degrees of freedom is multiplied
289: by four at each step.
290: As an approximation, Wilson suggested to keep at most a fixed number $m$
291: of the lowest-lying eigenstates of $H_L$ at each step.
292: For the Kondo problem, the error made at each step can be shown to be
293: of order $\Lambda^{-1/2} < 1$ \cite{wilson_rg}.
294:
295: \subsubsection{Numerical Procedure}
296:
297: \label{sec:NRG_procedure}
298:
299: The NRG method then proceeds as follows:
300: \begin{itemize}
301: \item[1)] Diagonalize $\mathbf{H}_L$ numerically, finding the $m$ lowest
302: eigenvalues and the corresponding eigenvectors.
303: \item[2)] Use the undercomplete similarity transformation $\mathbf{O}_L$ formed
304: by the $m$ eigenstates obtained in 1) to transform
305: all relevant operators on the $L$-site system to the new basis.
306: For example, $\bar{\mathbf{H}}_L = \mathbf{O}^\dagger \mathbf{H}_L
307: \mathbf{O}_L$ is a diagonal matrix of dimension $m$, but other
308: operators $\bar{\mathbf{A}}_L = \mathbf{O}^\dagger \mathbf{A}_L
309: \mathbf{O}_L$
310: will not, in general, be diagonal.
311: \item[3)] Form $\mathbf{H}_{L+1}$ from $\bar{\mathbf{H}}_L$ using the
312: recursion relation (\ref{eqn:NRG_recursion}), i.e., by adding a site
313: to the chain and constructing $\mathbf{H}_{L+1}$ in the expanded
314: product basis.
315: \item[4)] Repeat 1)-3), substituting $\mathbf{H}_{L+1}$ for
316: $\mathbf{H}_L$.
317:
318: \end{itemize}
319: This procedure is illustrated schematically in Fig.\ \ref{fig:NRG}.
320: The procedure can be started with the purely local term $H_0$
321: consisting of the impurity coupled to a single site, but in practice,
322: the relevant first step occurs when the dimension of $\mathbf{H}_L$ is
323: greater than $m$.
324: In step 1), one quarter of the eigenstates must be found at a general
325: step for the Kondo problem, so that typically ``complete
326: diagonalization'' algorithms as discussed in
327: Sec.~\ref{sec:ed_complete}
328: %srm such as the Householder transformation followed by the QL or QR
329: %algorithm
330: are used.
331: In general, the matrices to be diagonalized are block diagonal with
332: respect to the conserved quantum numbers of the system, such as number
333: of conduction electrons $N$ and the projection of the total spin
334: $S^z$.
335: As described in Sec.~\ref{sec::representation_manybodystates}, it is
336: important for numerical efficiency to separate these blocks.
337: Once this is done, the matrices to be diagonalized are not
338: particularly sparse.
339: In step 3), matrix elements of operators linking sites $L$ and $L+1$ such as
340: $f^\dagger_{L,\sigma} f_{L+1,\sigma}$ must be constructed in the
341: product basis $| i, s_L\rangle \equiv | i \rangle | s_{L+1} \rangle $,
342: with the first ket
343: representing the basis of $\bar{H}_L$, and the second the basis of the
344: added site (consisting of 4 states for the Kondo problem).
345: The expression for the matrix elements of $H_{L+1}$ is
346: \begin{eqnarray}
347: \langle i, s_L | H_{L+1} | i', s'_L \rangle & = &
348: \Lambda^{1/2} \; \langle i | i' \rangle \,
349: \langle s_{L+1} | s'_{L+1} \rangle \; E^L_i \nonumber \\
350: & & + \; (-1)^{N_{S_{L+1}}} \;
351: \langle i | f^\dagger_{L,\sigma}| i' \rangle \,
352: \langle s_{L+1} | f^{\phantom{\dagger}}_{L+1,\sigma}| s'_{L+1}\rangle
353: \\
354: & & + \; (-1)^{N_{S'_{L+1}}} \;
355: \langle i | f^{\phantom{\dagger}}_{L,\sigma}| i' \rangle \,
356: \langle s_{L+1} | f^{\dagger}_{L+1,\sigma}| s'_{L+1}\rangle \nonumber
357: \end{eqnarray}
358: where the $E^L_i$ are the eigenvalues of $H_L$, $i$ runs from 1 to
359: $m$, and $N_{S_L}$ is the number of electrons in state $S_L$.
360:
361: Note that at a particular point in the procedure, the range of
362: eigenvalues of $\tilde{H}_{L} = D_L H_L$ which accurately approximates
363: the spectrum of the infinite system is limited.
364: In particular, the accuracy breaks down due to the truncation at a
365: scale that is a multiple $\alpha$ of $D_L$ which depends on the value of
366: $\Lambda$ and the number of states kept $m$; typically $\alpha$ is of
367: order 10 when $m$ is of order 1000.
368: A lower limit is set by the energy scale $D_L$: eigenvalues below
369: $D_L$ are approximated more accurately at subsequent steps, i.e., for
370: $L'$ larger than $L$ and smaller $D_{L'}$.
371: Therefore, the non-rescaled eigenvalues $\tilde{E}_L$ in the range
372: $D_L \le \tilde{E}_L \le \alpha D_L$ can be accurately calculated at a
373: particular step.
374:
375: \begin{figure}
376: \includegraphics[width=0.6\textwidth]{NRG_figs/wilson.eps}
377: \caption{Schematic depiction of the NRG procedure.}
378: \label{fig:NRG}
379: \end{figure}
380:
381: \subsubsection{Renormalization Group Flow and Fixed Points}
382:
383: In order to understand the behavior of a system within the
384: renormalization group in general, one searches for fixed points of the
385: renormalization group transformation \cite{wilson_rg}, defined by
386: \begin{equation}
387: {\cal R}[H^*] = H^* \; .
388: \label{eqn:RG_fixed_point}
389: \end{equation}
390: In analytic variants of the renormalization group, the behavior of the
391: Hamiltonian $H$ is
392: generally parameterized by a small number of coupling constants.
393: Finding fixed points then amounts to finding stationary points in
394: the flow equations governing these coupling constants.
395: In the NRG, identifying fixed points is little more subtle: in
396: practice, when the first $p$ rescaled energy levels $E_p^L$
397: are independent of $L$ for a particular (appropriately chosen) range
398: of $L$, one identifies a fixed point.
399: Some insight into the physics of the problem is usually necessary
400: to choose appropriate ranges of $p$ and $L$.
401: Once fixed points are identified, the physicical behavior governed by
402: the fixed point is determined by the structure of the low-lying
403: eigenstates.
404: Note that such behavior is not guaranteed for a more general tight-binding
405: Hamiltonian.
406:
407: For the Kondo problem (and the SIAM in appropriate parameter
408: regimes), two fixed points can be clearly identified;
409: one associated with the behavior of model (\ref{eqn:kondo_chain})
410: at $J=0$ and one with its behavior at $J=\infty$.
411: In both cases, the behavior at the fixed point is easy to understand:
412: for $J=0$, the impurity is uncoupled to the conduction band and the
413: excitation energies are those of the non-interacting tight-binding
414: band extending from 0 to $L$.
415: For $J=\infty$, the impurity forms an infinitely tightly bound
416: singlet with site 0 of the tight-binding chain, effectively removing it
417: from the system.
418: The excitation energies relative to the ground state are therefore
419: those of a chain extending from 1 to $L$, i.e., the excitation
420: spectrum for $H_L(J=\infty)$ is the same as that for $H_{L-1}(J=0)$.
421: An additional complication is that the nature of the spectrum for
422: $H_L(J=0)$ depends on whether $L$ is even or odd; the asymptotic
423: values of the scaled excitation energies are different for even and
424: odd $L$ are different, even in the limit of large $L$.
425: This can be taken into account, however, by always applying a
426: sequence of two renormalization group transformations, i.e., by
427: replacing $\cal R$ with ${\cal R}^2$ in (\ref{eqn:RG_fixed_point}) to
428: determine the fixed points and then considering the odd $L$ and even
429: $L$ cases separately.
430: In fact, the crossover from the $J=0$ fixed point to the $J=\infty$
431: fixed point amounts to a reversal of the behavior for odd and even $L$
432: because the zeroth lattice site is effectively removed from the chain
433: at the strong coupling fixed point.
434: For details on the structure of the excitation energies, see
435: Refs.\ \cite{wilson_rg,hewson_book,krishna-murthy}.
436:
437: Numerically, one finds that for small $L$ and small but finite $J$,
438: the structure of the excitations is that of the $J=0$ fixed point,
439: whereas for large $L$ the structure is that of the $J=\infty$ fixed
440: point.
441: Stability analysis, which can be carried out analytically near the fixed
442: points, shows that the effective Hamiltonian for the weak-coupling
443: fixed point has a marginal operator, indicating that it is unstable,
444: while the effective Hamiltonian for the strong-coupling fixed point has no
445: relevant operators, indicating that it is stable
446: \cite{wilson_rg,krishna-murthy,hewson_book}, in agreement with
447: the behavior observed in the NRG.
448: The renormalization group flow from the unstable $J=0$ fixed point to
449: the stable $J=\infty$ fixed point is depicted in
450: Fig.\ \ref{fig:RG_flow}.
451:
452: \begin{figure}
453: \includegraphics[width=0.5\textwidth]{NRG_figs/RG_flow.eps}
454: \caption{Renormalization group flow diagram for the Kondo model.}
455: \label{fig:RG_flow}
456: \end{figure}
457:
458: \subsubsection{Calculation of Thermodynamic Properties}
459:
460: Thermodynamic quantities such as the specific heat or the impurity
461: susceptibility can be easily calculated within the NRG if the range of
462: validity of the excitation spectrum is taken into account.
463: Generally, one is interested in the impurity contribution to the
464: thermodynamic quantities, derived from the impurity free energy
465: $F_{\rm imp}(T) = -k_B T \ln Z / Z_c$, where $Z_c$ is the exactly
466: calculable partition function for the noninteracting conduction band.
467: If the entire eigenvalue spectrum were known, the partition function
468: would be given by $Z(T) = {\rm Tr} \exp (-\tilde{H}^K/k_B T)$.
469: However, at a particular stage of the NRG, what one can calculate is
470: the partition function for the truncated lattice
471: \begin{equation}
472: Z_L(T) \equiv {\rm Tr} e^{-\tilde{H}_L/k_B T}
473: = \sum_n e^{-\tilde{E}^L_n/k_B T}
474: = \sum_n e^{- D_L E^L_n/k_B T} \; .
475: \end{equation}
476: Evidently, $Z_L(T)$ can only be a good approximation for $Z(T)$ when
477: the temperature for a particular system size $T_L$ is chosen so that
478: $k_B T_L \ll \alpha D_L$, the largest
479: energy scale accurately described by $\tilde{H}_L$.
480: We previously argued that the minimum energy is set by $D_L$; the
481: error made in substituting $Z_L$ for $Z$ in calculating impurity
482: properties has been more rigously estimated to be $D_L/\Lambda k_B T$
483: in Ref.\ \cite{krishna-murthy}.
484: Therefore, the valid temperature range for a given $L$ is set by
485: \begin{equation}
486: \Lambda^{-1} \ll \frac{k_B T_L}{D_L} \ll \alpha \; ,
487: \end{equation}
488: where $\alpha$ depends on $m$ and $\Lambda$.
489: In practice, $k_B T_L \approx D_L$ is a reasonable choice.
490:
491: Experimentally interesting quantities are the impurity
492: specific heat
493: \begin{equation}
494: C_{\rm imp} = - T \frac{\partial^2}{\partial T^2} F_{\rm imp}(T)
495: \end{equation}
496: and the magnetic susceptibility at zero field due to the impurity
497: \begin{equation}
498: \chi_{\rm imp} = \frac{(g \mu_B)^2}{k_B T} \left [
499: \frac{{\rm Tr} (S_L^z)^2 \, e^{-\tilde{H}/k_B T}}{Z} -
500: \frac{{\rm Tr} (S_L^{z,c})^2 \, e^{-\tilde{H}_c/k_B T} }{Z_c}
501: \right ]
502: \end{equation}
503: where $S_L^z$ and $S_L^{z,c}$ are the $z$-components of the total spin
504: on the $L$-site chain with and without the impurity spin,
505: respectively.
506: The Hamiltonian $H_c$ is that of the noninteracting conduction band.
507: The high- and low-temperature limits, as well as the leading behavior
508: around these limits have been calculated analytically.
509: These limits can serve as a check of the accuracy of the NRG
510: calculations.
511: For a more complete depiction and discussion of results for
512: thermodynamic properties, see
513: Refs.\ \cite{wilson_rg,krishna-murthy, hewson_book}.
514:
515: \subsubsection{Dynamical Observables and Transport Properties}
516:
517: Dynamical properties, both at zero and at finite temperature, can also
518: be calculated within the NRG procedure.
519: To be concrete, we will discuss perhaps the most experimentally
520: interesting quantity, the impurity spectral function at zero
521: temperature (for simplicity):
522: \begin{equation}
523: A(\omega) = -\frac{1}{\pi} \,{\rm Im} \, G(\omega + i\eta) \; ,
524: \end{equation}
525: where
526: \begin{equation}
527: G(t) =
528: -i \langle \psi_0 |\, T \, d(t) d^\dagger(0) \, | \psi_0 \rangle
529: \end{equation}
530: is the retarded impurity Green function
531: [c.f.\ Eqs. (\ref{eqn:t_Green_fctn})-(\ref{eqn:lehmann_repn})].
532: For finite $L$, it is convenient to calculate the spectral weight
533: within the Lehmann representation
534: \begin{equation}
535: A_L(\omega) = \frac{1}{Z_L} \sum_{p}
536: \left | \langle p | d^\dagger_\sigma | 0 \rangle \right |^2
537: \delta(\omega - E_p + E_0)
538: \, + \,
539: \left | \langle 0 | d^\dagger_\sigma | p \rangle \right |^2
540: \delta(\omega + E_p - E_0) \; .
541: \label{eqn:lehmann_impurity}
542: \end{equation}
543: Since the excitations out of the ground state for $\tilde{H}_L$ are
544: well-represented in the energy range $D_L \le \omega \le \alpha D_L $,
545: as discussed previously, $A(\omega_L) \approx A(\omega)$ when
546: $\omega$ is chosen to be within this range.
547: In practice, a typical choice is $\omega = 2 \omega_L$ where
548: $\omega_L \equiv k_B T_L = D_L$.
549: One usually uses Eq.\ (\ref{eqn:lehmann_impurity}) directly to
550: calculate the spectrum, rather than the more sophisticated methods
551: such as the Krylov method outlined in Sec.\ \ref{sec:cont_fraction} or
552: the correction vector method outlined in
553: Sec.\ \ref{sec::ed_correction_vector} because all eigenstates within
554: the required
555: range of excitation energy are available within the NRG procedure; it
556: is then easy to calculate the poles and matrix elements within this
557: range.
558: Note that the result obtained is a set of positions and weights of
559: $\delta$-functions; in order to compare with continuous experimental
560: spectra, they must be broadened.
561: Typical choices for a broadening function are Gaussian or Logarithmic
562: Gaussian distributions of width $D_L$ \cite{sakai,costi_dmrg_book}.
563:
564: One interesting application of impurity problems comes about in the
565: context of the
566: Dynamical Mean Field Theory (DMFT), in which a quantum lattice model
567: such as the Hubbard model is treated in the limit of infinite
568: dimensions \cite{metzner-vollhardt,jarrell,georges_RMP}.
569: The problem can be reduced to that of a generalized SIAM interacting
570: with a bath or host.
571: The fully frequency-dependent impurity Green function of this
572: generalized SIAM must then be calculated in order to iterate a set of
573: self-consistent mean-field equations.
574: Various methods can be used to solve the impurity problem including
575: perturbation theory, exact diagonalization, quantum Monte Carlo,
576: dynamical DMRG (DDMRG), and the NRG \cite{georges_RMP,bulla_DMFT}.
577: Dynamical properties at finite temperature can be calculated
578: using similar considerations, as long as the additional energy scale
579: $k_B T$ is taken into account
580: \cite{costi-hewson91,costi-hewson92,costi-hewson93,costi_dmrg_book}.
581: In particular, the procedure outlined above can be used as long as
582: frequencies $\omega > k_B T$ are considered.
583: For $\omega < k_B T$, additional excitations at higher energies than
584: $D_L < \omega < \alpha D_L$ become important.
585: Frequencies in this range can then be handled using a {\sl smaller}
586: $L$ so that $\omega_L < K_B T$.
587: Since transport properties such as the resistivity $\rho(T)$ can be
588: formulated in terms of integrals over frequency of frequency-dependent
589: dynamical correlation functions \cite{costi-hewson93}, these methods
590: can also be used to calculate them.
591:
592: \subsection{Numerical RG for Quantum Lattice Problems}
593:
594: There are a number of quantum lattice problems that have Hamiltonians
595: whose structure is formally similar to the tight-binding
596: Hamiltonian for the Kondo model (\ref{eqn:kondo_chain}) such as the
597: Hubbard model (\ref{eqn:hubbard}) or the Heisenberg model
598: (\ref{eqn:heisenberg}) in one dimension.
599: While one could consider carrying out a variation of the NRG procedure on
600: these models in order to perform an approximate exact diagonalization
601: on a finite lattice, the physically interesting case is the one in
602: which the couplings between nearest-neighbor sites are all equal.
603: This amounts to setting $\Lambda=1$ in the Kondo case, the point at
604: which the convergence of the NRG method breaks down completely because
605: the identification of the length $L$ with the energy scale is lost.
606: Nevertheless, adaptations of the NRG procedure were applied to small
607: Hubbard chains \cite{bray-chui}, obtaining an error
608: in the ground-state energy of approximately 10\% after 4
609: renormalization group steps.
610: Later work on the spin-1 Heisenberg chain obtained an error of
611: approximately 3\% in the ground-state energy of the $L=18$ chain
612: \cite{xiang-gehring}.
613: It is important to note that such calculations are variational so that
614: quantities which characterize the long distance behavior such as
615: correlation functions which depend on the wave function can have much
616: larger errors than the error in the energy.
617: Finally, an adaptation of the NRG procedure to a two-dimensional
618: noninteracting electron gas was used to study the Anderson localization
619: problem in two dimensions \cite{lee}.
620: The result obtained was that the system undergoes a
621: localization-delocalization transition as a function of disorder
622: strength, a result later discovered to be incorrect: there is no
623: transition because two is the lower critical dimension
624: \cite{lee_fisher,bigfour}.
625:
626: \subsection{Numerical RG for a Noninteracting Particle}
627: \label{sec::nrg_noninteracting}
628: More insight into the breakdown of the NRG procedure for
629: one-dimensional lattice problems with non-decaying couplings can be gained by
630: applying a variant of the procedure to the problem of a single
631: particle on a tight-binding chain, Eq.~(\ref{eqn:tb}).
632: We consider the Hamiltonian in the formulation
633: \begin{equation}
634: H = -\sum_{\ell=1}^{L-1} \left ( ~
635: | \ell \rangle \langle \ell+1 | + | \ell+1 \rangle \langle \ell | ~\right )
636: + 2 \sum_{\ell=1}^L |\ell \rangle \langle \ell | \; ,
637: \label{eqn:nonint_tight_bind}
638: \end{equation}
639: where that state $|i\rangle$ represents an orbital localized on site
640: $\ell$.
641: In a matrix representation, this Hamiltonian is tridiagonal and
642: is equivalent to a discretized second derivative operator,
643: $-\partial^2/\partial x^2$.
644: Note that Hamiltonian (\ref{eqn:nonint_tight_bind}) does not include a
645: nonzero matrix element between sites 1 and $L$, so that fixed boundary
646: conditions have been applied to the chain, i.e., the wave function is
647: required to vanish at the ends.
648: We modify the Wilson NRG procedure slightly to take into account that
649: we are treating a simpler, noninteracting system.
650: There are two significant changes: first, we put together
651: two equal-sized systems (called ``blocks'') rather than just adding a
652: site because the dimension of the Hilbert space grows more slowly than
653: in an interacting system: linearly rather than exponentially with the
654: length.
655: Second, the mechanics of putting two blocks together is simpler in the
656: interacting system.
657:
658: In terms of the procedure outlined in Sec.\ \ref{sec:NRG_procedure}, the
659: diagonalization of the Hamiltonian, step 1), and the transformation
660: of the relevant operators using $\mathbf{O}_L$, a matrix whose columns
661: are the $m$ eigenvectors of $\mathbf{H}_L$ with the lowest
662: eigenvalues, [step 2)] are carried out as before.
663: The procedure is somewhat simplified because of the less complicated
664: structure of Hamiltonian (\ref{eqn:nonint_tight_bind}); for example,
665: there are no $S^z$ and $N$ quantum numbers which decouple sectors of
666: $\mathbf{H}$.
667: The operators to be transformed are the Hamiltonian for a block
668: $\bar{\mathbf{H}}_L = \mathbf{O}^\dagger_L \mathbf{H}_L \mathbf{O}_L$
669: which is diagonal and $\bar{\mathbf{T}}_L = \mathbf{O}^\dagger_L
670: \mathbf{H}_L \mathbf{O}_L$, where $\mathbf{T}_L$ represents the
671: connection between the blocks.
672: The procedure is conveniently started at $L=1$ for which $\mathbf{H}_1
673: = 2$ and $\mathbf{T}_L=-1$ can be represented as $1\times 1$
674: matrices.
675: Matrix representations of a system of size $2L$ are then formed [step 3)]
676: as
677: \begin{eqnarray}
678: \mathbf{H}_{2L}=
679: \left(
680: \begin{array}{cc}
681: \bar{\mathbf{H}}_L & \bar{\mathbf{T}}_L \\
682: \bar{\mathbf{T}}^\dagger_L & \bar{\mathbf{H}}_L \\
683: \end{array}
684: \right)
685: \label{eqnH_2L}
686: \end{eqnarray}
687: and
688: \begin{eqnarray}
689: \mathbf{T}_{2L}=
690: \left(
691: \begin{array}{cc}
692: 0 & 0 \\
693: \bar{\mathbf{T}}_L & 0 \\
694: \end{array}
695: \right) \; .
696: \label{eqnT_2L}
697: \end{eqnarray}
698: The procedure is then iterated [step 4)], doubling the size of the
699: system at each step.
700: The matrix to be diagonalized, $\mathbf{H}_{2L}$, is of size $2m\times
701: 2m$ at a general step, and the truncation is to an $m\times m$ matrix,
702: i.e., one half of the Hilbert space is discarded.
703: Note that the procedure is exact for the first few steps, as long as
704: $m \le L$.
705:
706: \begin{table}
707: \caption{ Lowest energies after 10 blocking
708: transformations for the
709: noninteracting single particle on a 1-D chain with fixed boundary
710: conditions, keeping up to $m=8$ states.}
711: %\begin{center}
712: \begin{tabular}{|c|c|c|c|}
713: \hline
714: {$\;$} & {Exact} & {Wilson} & {Fixed-Free} \\
715: \hline
716: $\ E_0\ $
717: & $\; 2.3508\times 10^{-6}\; $
718: & $\; 1.9207 \times 10^{-2} \; $
719: & $\; 2.3508 \times 10^{-6} \;$
720: \\
721: $\ E_1\ $
722: & 9.4032$\; \times 10^{-6}\; $
723: & 1.9209$\; \times 10^{-2}\;$
724: & $9.4032\times 10^{-6}$
725: \\
726: $\ E_2\ $
727: & $2.1157\times 10^{-5}$
728: & $1.9214\times 10^{-2}$
729: & $2.1157\times 10^{-5}$
730: \\
731: $\ E_3\ $
732: & $3.7613\times 10^{-5}$
733: & $1.9217\times 10^{-2}$
734: & $3.7613\times 10^{-5}$
735: \\
736: \hline
737: \end{tabular}
738: %\end{center}
739: %\vspace*{-0.4cm}
740: \label{tab:box_results}
741: \end{table}
742:
743: As shown in the first two columns of Table \ref{tab:box_results},
744: the accuracy of the eigenvalues obtained breaks completely after a
745: moderate number of steps.
746: The failure of the NRG for this simple problem as well as the reason
747: for the failure was pointed out by
748: Wilson in 1986 \cite{wilson_seminar}.
749: For a particular $L$, the eigenfunctions have the form
750: \begin{equation}
751: \psi^L_n(\ell) \propto \sin \left (
752: \frac {n \pi \ell}{L+1} \right ) \; ,
753: \hspace*{2cm} n=1,\ldots,L
754: \end{equation}
755: because fixed boundary conditions have been applied to the system.
756: In one iteration of the NRG procedure, a linear combination of the
757: $\psi^L_n(\ell)$ for $n \le m$ are used to form an approximation to
758: $\psi^{2L}_n(\ell)$.
759: As depicted in Fig.\ \ref{fig:pboxwf}, the boundary conditions that
760: are applied to $\mathbf{H}_L$ lead to a non-smooth ``dip'' in the
761: wave function in the middle of the block of size $2L$ when the NRG
762: transformation is carried out.
763: Clearly, this dip can only be removed by forming a linear combination
764: of almost all of the $\psi^L_n(\ell)$.
765: The lesson that is learned, then, is that how the
766: boundaries of the blocks are treated in the NRG procedure is crucial:
767: a more general treatment is necessary to formulate a numerically
768: accurate real-space NRG procedure for short-ranged quantum lattice
769: models.
770:
771: \begin{figure}
772: \includegraphics[width=0.5\textwidth]{NRG_figs/pboxwf.eps}
773: \caption{The lowest eigenstates of two 8-site blocks (solid circles) and a
774: 16-site block (open squares) for the one-dimensional tight--binding
775: model with fixed boundary conditions. }
776: \label{fig:pboxwf}
777: \end{figure}
778: