Equilibrium dynamics of the sub-Ohmic spin-boson model under bias
Zheng Da-Chuan, Tong Ning-Hua
Department of Physics, Renmin University of China, Beijing 100872, China

 

† Corresponding author. E-mail: nhtong@ruc.edu.cn

Abstract
Abstract

Using the bosonic numerical renormalization group method, we studied the equilibrium dynamical correlation function C(ω) of the spin operator σz for the biased sub-Ohmic spin-boson model. The small-ω behavior is found to be universal and independent of the bias ε and the coupling strength α (except at the quantum critical point and ε = 0). Our NRG data also show for a wide range of parameters, including the biased strong coupling regime ( and ), supporting the general validity of the Shiba relation. Close to the quantum critical point , the dependence of C(ω) on α and ε is understood in terms of the competition between ε and the crossover energy scale of the unbiased case. C(ω) is stable with respect to ε for . For , it is suppressed by ε in the low frequency regime. We establish that holds for all sub-Ohmic regime , with for and for . The variation of C(ω) with α and ε is summarized into a crossover phase diagram on the αε plane.

1. Introduction

The spin-boson model (SBM) is one of the simplest models to describe a quantum two-level system coupled to the environmental noise.[1,2] It has realizations in various fields of physics, including the superconducting qubit,[3] mesoscopic metal ring penetrated by an Aharonov–Bohm flux,[4] ultraslow glass dynamics,[5] heavy fermion metals,[6] and nanomechanical oscillators,[7] etc. In this model, the environmental noise is represented by a bath of harmonic oscillators with the low frequency spectral function characterized by . Here α is the coupling constant and s is a dimensionless parameter that discriminates the Ohmic (s = 1), super-Ohmic ( ), and sub-Ohmic ( ) bath.

The SBM has been studied in detail for the Ohmic and weak-coupling regime to understand the quantum decoherence and dissipation, which are the key issues for making a long-lived quantum bit,[3] chemical reaction,[8] as well as in the optical absorption in a quantum dot.[9] The SBM without a bias field also contains non-trivial quantum phase transitions (QPTs) between a weak-coupling delocalized phase and a strong-coupling localized phase for . It is an established result that a Kosterlitz–Thouless-type QPT exists for s = 1 and the ground state is always delocalized for .[1,2] In the sub-Ohmic regime ,[1,2,10] numerical methods such as numerical renormalization group (NRG),[1113] quantum Monte Carlo (QMC),[14] and exact diagonalization[15] show that a continuous QPT occurs at certain critical coupling , at which both the static and the dynamic quantities show critical behavior.[1625]

For the study of both decoherence and QPT, the equilibrium dynamical correlation function is an important quantity. Most of the previous studies concentrate on the case of zero bias and weak coupling, i.e., ε = 0 and . In this regime, the correlation function C(ω) (to be defined below) shows a power law behaviour in small frequency and critical behavior in the intermediate frequency . Here ω0 is a non-universal high energy scale and is the crossover energy scale between the delocalized fixed point and the quantum critical fixed point.

The situation with a finite bias , however, has received less attention. For , the parity symmetry of the unbiased SBM is broken and hence the localized–delocalized QPT of the sub-Ohmic symmetric SBM no longer exists. Instead, as the coupling strength increases, the ground state changes smoothly from the biased weak-coupling delocalized-like state to the strong-coupling localized-like state with broken symmetry. The strong coupling regime is difficult to describe for the perturbation-based theory.[18] For numerical approaches such as NRG[11,12] and exact diagonalaization,[15] a finite bias increases the number of bosons in the ground state and makes an accurate calculation more difficult. As a result, systematic study of the dynamical correlation function for the sub-Ohmic bath under a finite bias, especially in the strong coupling regime , is still lacking.

In this paper, we use NRG to study the equilibrium state dynamical correlation function of the sub-Ohmic SBM with a finite bias field , with emphasis on the strong-coupling regime . By carefully extrapolating the boson-state truncation to infinity, we find that except for the exact QPT point, C(ω) always obeys behavior in the small frequency limit, irrespective of the values of ε and α. In the vicinity of the quantum critical point (QCP), C(ω) is characterized by two different power law regimes: in the low frequency regime and in the intermediate frequency regime , with ω0 being a non-universal high energy scale. The crossover frequency is tuned by ε. For weak bias , and C(ω) is not changed significantly from the zero bias case. For strong bias , and C(ω) is suppressed in . The behavior of is understood in terms of the competition between the unbiased crossover scale and ε, and we have . We obtained the critical exponent θ as a function of s. The crossover bias is thus determined as . We finally summarize the behavior of C(ω) in a crossover phase diagram.

This paper is organized as the following. In Section 2 we introduce SBM and the formalism we used to calculate the equilibrium correlation function with thev bosonic NRG method. Section 3 presents the results from our NRG study. A conclusion is given in Section 4.

2. Model and method

The Hamiltonian of SBM reads

The first two terms describe a two level system with bias ε and tunnelling strength . The bosonic bath is described by the third term, where ωi is the frequency of the i-th boson mode. In the last term, the two level system is coupled to the bosonic bath through σz and the boson displacement operator. Various experimental realizations of this Hamiltonian have been proposed with tunable parameters ε and .[4,2628] The influence of the bath on the two level system is characterized by the spectral function
In this paper, we use a power law form of with a hard cut-off at ,
We set as the unit of energy and fix to study the dependence of equilibrium dynamics on ε and α for a sub-Ohmic bath.

At ε = 0, equation (1) is invariant under the combined boson and spin parity transformation and . For the sub-Ohmic ( ) and the Ohmic ( ) baths, a spontaneous breaking of this symmetry may occur in the regime and the system enters the localized phase, in which the quantum system is trapped to one of the two states and the local bosons have a finite displacement.[1015] This is the delocalize–localize quantum phase transition of SBM. With a finite bias , the above parity symmetry is broken from the outset and the phase transition no longer exists. As α increases, the ground state crosses over smoothly from the weak-coupling to the strong-coupling biased states.

In this paper, we focus on the dynamical correlation function defined as

where and is the anti-commutator of and . For a non-degenerate ground state, the Lehman representation of the correlation function at is written as
Here and En are the n-th eigen state and energy of the Hamiltonian, respectively. C(ω) has the general form , where and is the ground state. It is an even function of ω and fulfils the sum rule

The bosonic NRG method is regarded as one of the most accurate numerical techniques for studying SBM due to its non-perturbative nature and the applicability in the whole range of parameters.[11,12,29,30] The success of NRG relies on the energy scale separation due to the logarithmic discretization, and on the RG transformation which is carried out by iterative diagonalization in each energy shell. It is technically composed of three steps: logarithmic discretization, transforming the Hamiltonian into a semi-infinite chain, and the iterative diagonalization. In general, the errors in the NRG calculation come from two sources. One is the approximation of using one bath mode to represent each energy shell, which is controlled by the logarithmic discretization parameter . The other is the truncation of the energy spectrum after each diagonalization to overcome the exponential increase of the Hilbert space, which is controlled by the number of kept states Ms. For the bosonic NRG, an additional source of error is the truncation of infinite dimensional Hilbert space of each boson mode into states on the occupation basis. Exact results are obtained in the limit , , and .

In practice, however, one cannot do the calculation directly at the above limit and extrapolating NRG data to the above limit could be difficult. In the strong coupling regime with a bias, the ground state has a large boson number and hence requires a large for an accurate calculation. A large , however, will increase the truncation error because in the NRG calculation only the lowest fraction of the eigen states of the previous energy shell are kept for constructing the Hamiltonian of the lower energy shell. To keep the accuracy, one has to use a larger to increase the energy separation of successive energy shells. This in turn will lead to a larger discretization error of the bath. Fortunately, we found that a larger only induces quantitative changes to the NRG results and the qualitative conclusion can still be obtained reliably. Therefore, using a large , a moderate , and , we can obtain qualitatively correct results for the biased SBM in the strong coupling regime. Combining the knowledge gained from the finite size scaling method of ,[31,32] we checked that our final conclusions are stable when extrapolated to the exact limit. To calculate C(ω), we use the patching method of Bulla[33] with which the sum rule Eq. (6) is fulfilled reasonably well. The discrete δ peaks obtained are broadened with a Gaussian function on the logarithmic scale,[33]

We choose the broadening parameter b = 0.7 for and b = 1.2 for .

3. Results

In this work, we study the dynamical correlation function C(ω) for all the range of α and . As ε tends to zero, the strong coupling phase of is continuously connected to the localized state at with spontaneously broken symmetry. At ε = 0 and , the localized state without symmetry breaking is special in that the ground state has a two-fold degeneracy and it is unstable in the thermodynamical limit.[34] In this paper we hence confine our study to the symmetry-broken phase (either spontaneously or by a finite bias) and do not study that special case.

3.1. -behavior in the small frequency regime

For the Ohmic (s = 1) and the super-Ohmic ( ) baths, the small frequency behavior of C(ω) in the delocalized phase obeys the Shiba relation,[35] which in our notation reads[12]

Here is the local spin susceptibility. The proof in Ref. [35] applies also to the case with a finite bias, but not to the localized phase at for the Ohmic bath. This exact relation was used to test the quality of various approximate results for SBM.[11,12,19,3639]

There have been attempts to generalize the Shiba relation to (i) sub-Ohmic bath, (ii) with finite bias, and (iii) strong coupling regime . Such a generalized relation, if it exists, would imply a universal long time behavior for any s values, coupling strength α (except at the critical point and ε = 0), and spin parameters and ε. Up to now these activities received partial success only. Florens and others[40] proved Eq. (8) for the sub-Ohmic bath, based on an exact relation from the perturbation theory in Majorana representation. However, their proof applies only to delocalized phase at the symmetry point ε = 0. The numerical results from the approximate perturbation theory built on a unitary transformation fulfils Eq. (8) exactly for the sub-Ohmic bath with[41,42] or without[18] bias in the weak to intermediate coupling. In the strong coupling regime the fulfilment is good but not exact. This leaves the question open whether the Shiba relation holds in the biased strong coupling regime of the sub-Ohmic SBM.

As stated above, the inherent truncation errors of NRG hinder it from the quantitative confirmation/falsification of the generalized Shiba relation. However, qualitatively, it is possible to check the factors on the right-hand side of Eq. (8) each for a time. In this section, we will show that holds for the sub-Ohmic SBM with general α and ε. In Fig. 6, data are presented to further support that for general α and ε values.

In Fig. 1, we first present the -dependence of C(ω) obtained from NRG under bias, in order to gauge the choice of in our study. The energy flows (Fig. 1(a)) and C(ω) (Fig. 1(b)) are shown for a sub-Ohmic bath s = 0.8 in the strong coupling regime , with a finite bias . We use a series of values from to with and . In Fig. 1(a), for a given , the excitation energies (i = 1, 2, 3) first flow to an unstable fixed point (with ) in the small N regime. After a crossover , they flow to the stable fixed point (with ) in the large N limit. We found that the unstable fixed point in the small N regime is the same as the weak-coupling fixed point of the SBM in the non-bias case (dashed lines obtained using α = 0 and ε = 0). With increasing geometrically, the crossover increases linearly, showing that the associated energy scale decreases to zero as a negative power of . This supports that the unstable fixed point (dashed lines in Fig. 1(a)) will extend to infinitely large N in the limit and it is the true fixed point of the biased SBM. In this biased fixed point, ε flows to infinity and the spin is effectively decoupled from the bath, leading to the same excitation levels as the free boson chain.

Fig. 1. (color online) NRG flow diagrams and C(ω) at s = 0.8, Δ = 0.1, , and . (a) Flow diagrams obtained using , 18, and 40 (symbols with eye-guiding lines, from left to right). The dashed lines are the flow of α = 0 and ε = 0. (b) Dynamical correlation function C(ω) for , 18, and 40 (solid lines, from bottom to top), and from extrapolation (dashed lines). Fitting the small ω regime, we obtain the exponents for , and for . The peak is not shown. NRG parameters are Λ = 4.0 and .

Figure 1(b) shows the corresponding C(ω), which has two different power-law regimes: in the low frequency regime , and in the high frequency regime . NRG data for s = 0.8 gives and . With increasing , the crossover frequency decreases to zero as a negative power of , proving that is the correct result of the biased SBM (dashed line in Fig. 1(b)). Our results for other s values in agree with within an error of 2% (not shown).

For , the NRG results converge with much more difficulty with increasing . In Fig. 2, we show the strong coupling data ( ) for s = 0.3. Both the energy flow and C(ω) have -dependent crossover scales. With increasing to 60, a clear trend can be seen that the -converged levels flow towards the free boson energy levels (Fig. 2(a)). For both and , the section of C(ω) with ωs behavior extends to smaller ω with increasing (Fig. 2(b)). The shift of C(ω) with has an apparent scaling form and an analysis for could be carried out to extract the correct exponent, as done in Ref. [31]. Here in Fig. 2(b), we only mark out the expected asymptotic ωs line for guiding the eye. With this understanding of the -dependence of C(ω), in the rest of this paper, we only show the physically correct results obtained using sufficiently large .

Fig. 2. (color online) NRG flow diagrams and C(ω) at s = 0.3, Δ = 0.1, . (solid lines) and (dashed lines). (a) Flow diagrams obtained using , 27, and 60 (from left to right). The horizontal dashed lines are the flow of α = 0 and ε = 0. (b) C(ω) for , 27, and 60 (from bottom to top). The functions and are marked by the dashed straight lines for guiding the eye. The peak is not shown. NRG parameters are Λ = 4.0 and .

In Fig. 3, we investigate the influence of ε on C(ω) for α values ranging from weak coupling (Fig. 3(a)) to strong coupling (Fig. 3(b)) for s = 0.8 and . It is seen that in the small frequency limit for all parameters.In Fig. 3(a), for the small coupling α = 0.1 (dash-dotted lines), C(ω) does not change with ε up to and the three curves for ε = 0, 10−3, and 10−5 overlap. While for a moderate coupling α = 0.4 (dashed lines), a slight downward shift is observed between and and the lines for ε = 0 and 10−5 overlap. C(ω) is most sensitive to ε when α is close to . The same trend is observed in Fig. 3(b) for the strong coupling regime.

Fig. 3. (color online) C(ω) for various α and ε at s = 0.8, Δ = 0.1. (a) : α = 0.48 (solid lines), 0.4 (dashed lines), and 0.1 (dash-dotted lines). (b) : α = 0.49 (solid lines), 0.57 (dashed lines), and 0.7 (dash-dotted lines). For each α value, from top to bottom are curves for ε = 0, 10−5, and 10−3. All curves show behavior in the small ω limit. NRG parameters are Λ = 4, , and .

The universal behavior observed so far in Figs. 1, 2, and 3 has a natural understanding. When the parity symmetry is already broken by a finite bias, the ground state of SBM can be tuned continuously on the αε plane, going from a delocalized state at , , through a half circle in a finite ε region, to the symmetry-spontaneously-broken state at , , without passing the critical point. Therefore, the ground state has the same nature and no qualitative change is expected in the small ω limit of C(ω). Indeed, our NRG results for different s, α, and ε confirm that is a universal feature of SBM.

In the strong coupling case shown in Fig. 3(b), the same tendency is observed, i.e., with increasing ε, C(ω) shifts downwards and the most prominent change occurs near . The suppression of C(ω) can be understood as the weight transfer: with increasing ε, increases and so does the weight of the zero-frequency δ peak. Due to the sum rule, decreases uniformly. Comparing figs. 3(a) and 3(b), one sees that a broad peak forms for and .

Besides the universal low frequency behavior of C(ω) that we focus on in this paper, the high frequency structure in C(ω) for the deep sub-Ohmic SBM ( ) has also attracted a great deal of attention recently.[17,18,41] For small s values, a peak at the renormalized tunnelling is observed in the strong coupling regime. It persists even in the localized phase . For α close to and small ε, another pronounced peak is observed at the crossover scale . C(ω) thus has a two-peak structure for small s and close to the critical point. Here, our C(ω) for s = 0.3 and a weak bias does have the two-peak structure, as shown by the dashed lines in Fig. 2(b). The high frequency peak is weak due to the over broadening of the log-Gaussian function. Decreasing the broadening parameter b from 1.2 to 0.8, the high frequency peak gets sharper and more pronounced, being consistent with previous results.[17,18,40]

3.2. Equilibrium dynamics near QCP

In the above section, we established the universal ωs-behaviour of C(ω) in the full parameter space of SBM. In this section, we focus on the parameter regime near the QCP and study the critical properties of C(ω) under a bias. For the unbiased SBM, is the only energy scale that controls the crossover between the stable fixed points (localized and delocalized fixed points) and the critical fixed point. Here ν is the critical exponent of the correlation length and z = 1 is the dynamical critical exponent. plays an important role in the temperature dependence of physical quantities close to QCP. At zero temperature, it also appears in the dynamical correlation function in the delocalized phase: for and for , and . is the critical behavior which holds for arbitrarily small frequency at the exact critical point , and is confined to away from the critical point. As α approaches , tends to zero and the behavior will be recovered in the full frequency range.

For the biased case, is another energy scale that influences the crossover between different behaviors of C(ω). Our NRG results for C(ω) at different α and ε can be understood in terms of the competition between ε and . Due to the difficulty of convergence for , here we show NRG data for s = 0.8, representing a typical case for . Our conclusion also applies to , as will be discussed below.

In Fig. 4 and Fig. 5, we show the flow diagrams (Fig. 4) and C(ω) (Fig. 5) for (a), (b), and (c), respectively. For each α, ε varies from zero to 10−3. The purpose is to observe the influence of ε in the delocalized, localzied, and critical phases. Figures 4(a) and 5(a) are for . At ε = 0, the energy flow in Fig. 4(a) has a crossover at around (solid circle), from the critical fixed point to the weak-coupling one. As ε increases from zero, the flow does not change for and the crossover (empty squares) begins to decreases only for . The corresponding evolution of C(ω) is shown in Fig. 5(a). For , for and for . The crossover scale of the symmetric SBM is given by . With increasing ε, the crossover frequency increases and occurs only when , corresponding to the occurrence of in the flow diagram. A crossover scale can be defined as such that for , becomes significantly larger than , or equivalently, . For , C(ω) stays same as the symmetric case (ε = 0) and . For , C(ω) is suppressed in the low frequency regime and is set by ε.

Fig. 4. (color online) (a)–(c) The flow diagrams for three different α values. For each α, from right to left ε = 0, 10−7, 10−6, 10−5, 10−4, and 10−3. The empty squares make out the crossover for the case . The solid circle marks out the crossover scale in the unbiased case . In panel (c), . NRG parameters are , , and .
Fig. 5. (color online) (a)–(c): C(ω) for the same α values. For each α, from top to bottom, ε = 0, 10−7, 10−6, 10−5, 10−4, and 10−3. NRG parameters are the same as Fig. 4.

Figures 4(b) and 5(b) show the influence of ε in the localized phase. In Fig. 4(b), the excitation energy level at ε = 0 flows from the critical fixed point towards a two-fold degenerate fixed point, with the critical-to-localize crossover around (solid circle). Now we study the change of under bias (empty squares). With a vanishingly small ε, the degeneracy is lifted and the excited energy level has an upturn at an arbitrarily large , showing a crossover from the localized symmetric state to the biased state. We find that the crossover energy scale . With further increasing ε, the upturn moves to the left and for , occurs, which means that is now the crossover from the critical fixed point to the biased fixed point. NRG data give the critical-to-biased crossover . The critical exponent θ will be discussed with Fig. 5(c).

In Fig. 5(b), the evolution of C(ω) with ε is shown, which looks similar to the case . That is, C(ω) does not change much with ε for and begins to be suppressed for . It is noted that for a given , the degenerate-to-biased crossover at in the energy flow has no correspondence in C(ω): the latter has a perfect ωs behavior at . The crossover in the energy flow will only show up in the temperature dependence of C(ω).

Figures 4(c) and 5(c) show the influence of ε on the flow and C(ω) in the critical regime . Since at this point, ε is the only energy scale that controls the critical-to-biased crossover in the energy flow and C(ω). Especially, Fig. 5(c) shows that for and for . We define a critical exponent θ as

For s = 0.8, the fitted exponent from Fig. 5(c) is . Note that the same dependence of on ε applies to the and cases in the regime , which are shown in panels (a) and (b) of figs. 4 and 5.

The NRG result in Fig. 5(c) shows that for , with an ε-independent factor c. Combining this observation with the assumption that the Shiba relation Eq. (8) holds at and (which will be discussed in Fig. 6 below), we can derive θ by equating the small and the large frequency expression at , giving

Employing the critical behavior , one obtains , giving . From the exact expression δ = 3 for and for , we obtain
For s = 0.8, this expression gives θ = 1.111, which agrees well with the NRG result 1.12.

Fig. 6. (color online) Checking the Shiba relation at s = 0.8 and Δ = 0.1. C(ω) (empty symbols) is compared to (solid symbols) for various α and ε values. Here . NRG parameters are , , and .

Having shown that ( ) for general α and ε values in the previous section, we now check the Shiba relation Eq. (8) at s = 0.8 using a fixed small frequency . Figure 6 shows C(ω) and the right-hand side of Eq. (8) as functions of ε, for the same α values as in figs. 4 and 5. We find qualitative agreement between them for a wide range of ε. For ( ), C(ω) shows the crossover from critical (power law dependence on ε) for to the localized-like (delocalized-like) behavior (being constant) for , as expected. For , a power-law dependence is observed, with the fitted exponent −1.88 (for C(ω)) and −1.84 (for ), respectively, in reasonable agreement with the exact . It is notable that although C(ω) and have more than 5 decades of variations in the range , their ratio does not change much. This result supports the relation .

The ratio also depends weakly on ω due to the slight inaccuracy in the exponent of NRG-produced C(ω). The uniform deviation from Eq. (8) observed here is more likely due to the error of NRG data than due to the invalidity of the Shiba relation. After all, a factor of 3 is a reasonable level of error in the NRG calculation of dynamical quantities,[11,12] considering the logarithmic error, truncation errors, as well as the approximation used to calculate C(ω) and χ.[33] It is expected that the agreement can be improved if we increase Ms and Nb, and extrapolate to unity, which will not be pursued here. The good agreement of the NRG value of the exponent θ and Eq. (11) is also consistent with , since equation (11) is derived from this assumption. For s = 0.3, it is more difficult to check the Shiba relation quantitatively due to much stronger boson state truncation error. As shown in Fig. 2(b), using as large as 60, one can obtain the expected ωs behavior only in a narrow frequency window and for certain ε values. For those parameter regimes and frequencies where ωs behavior can be obtained, NRG results give a ratio for the delocalized, critical, and the localized phases, much larger than the s = 0.8 case. This shows the difficulty of a quantitatively accurate NRG study for the deep sub-Ohmic SBM.

In Fig. 7(a), we fix and plot C(ω) for different α values. For , C(ω) increases with increasing α and the height of the peak reaches a maximum at certain . With further increasing α, C(ω) begins to decrease uniformly, with the peak height suppressed. In Fig. 7(b), we show the maximum of C(ω) as functions of α for various εʼs. From this figure we can extract the α value at which the height of the peak in C(ω) reaches the maximum.

Fig. 7. (color online) (a) C(ω) at and for a series of αʼs. From top to bottom, α = 0.39, 0.385, 0.38, 0.375, and 0.37 (solid lines), and α = 0.395, 0.40, 0.405, and 0.41 (dashed lines). (b) The maximum value of C(ω) as functions of α for a series of ε. From top to bottom, ε = 0.0, , , , , , and . s = 0.8, Δ = 0.1. NRG parameters are Λ = 2.0, , and .

The phase diagram on the αε plane shown in Fig. 8 summarizes our results for C(ω) near the QCP. Close to and with very small ε, three regions I, II, and III of different nature are separated by a crossover line (squares with eye-guiding line). Regions I and III are continuously connected to the delocalized ( ) and localized phase ( ) of the symmetric SBM, respectively. They are characterized by . In region I, with χ being the magnetic susceptibility. In region III, close to the saturate value. For each α, the crossover in Fig. 8 is defined as the ε value at which the maximum of C(ω) decreases to 90% of its value at ε = 0.

Fig. 8. (color online) Phase diagram of sub-Ohmic SBM near QCP extracted from C(ω). Regions I, II, and III are the delocalized, quantum critical, and the localized phases, respectively. They are separated by a crossover line (solid squares with eye-guiding line) at which the peak height of C(ω) drops to 90% of its ε = 0 value. The peak height of C(ω) reaches maximum at the dashed line with empty circles. Inset: log-log plot of versus for (solid squares) and (solid circles). The fitted exponents are 1.989 and 1.985, respectively.

In region II, vanishes near and becomes the only characteristic energy scale. In this regime, the magnetization shows the quantum critical behavior as , with δ being a critical exponent. The empty circles with eye-guiding line marks out the parameter at which C(ω) has a highest peak. Both the crossover line and the peak position can be extracted from the data in Fig. 7(b). It is observed that the location of the maximum peak resembles that of the maximum of entanglement entropy.[21] This is not a pure coincident since the crossover peak of C(ω) reflects a strong fluctuation near the critical point and it is therefore naturally related to the entanglement maximum.

In the inset of Fig. 8, we fit the crossover line close to the QCP in a power law

with the obtained critical exponent η = 1.99 for s = 0.8. Note that the asymmetry of in and regimes means different pre-factors c. Besides using C(ω) here, the critical behavior of can also be determined by the scaling form of obtained from the scaling analysis of the NRG data.[32] These two approaches give consistent results. For the second approach, the scaling analysis of NRG data gives (see Eqs. (31), (33), (35), (36) in Ref. [32])
Here and is a two-variable scaling function. By comparing the magnitude of the two variables in Eq. (13), one obtains , giving . Here, β is the critical exponent of the order parameter. It is known that for , β = 1/2 and δ = 3. For , β is a function of s whose explicit expression is unknown yet, and .[13,14,31,43,44] The NRG result is in reasonable agreement with the fitted exponent η = 1.99 in the inset of Fig. 8, showing the consistency of the static and dynamical approaches.

For , NRG calculation is hindered by the slow convergence with and it is difficult to obtain the quantitative data. However, we note that the scaling from Eq. (13) is obtained from the NRG calculation combined with the scaling analysis for the full sub-Ohmic regime . It is stronger than the scaling ansatz of free energy which applies only to the regime .[43,44] The two exponents introduced above, θ and η, are not independent. We have . This gives , meaning . This is a relation independent of the validity of hyperscaling relation.

4. Conclusion

In summary, we use the bosonic NRG method to study the equilibrium dynamics of the sub-Ohmic SBM under a bias. We found that the small ω behavior holds for any bias ε and coupling strength α except exactly at the critical point , ε = 0, where . Our results strongly support that in all parameter regimes, including the biased and strong coupling regime. This is in favour of the validity of the generalized Shiba relation in this regime. Close to the QCP, the competition between the energy scale and bias ε determines a crossover scale which separate the weak-biased regime ( ) from a strong-biased regime ( ). Near the QCP, with . These results are summarized in the crossover phase diagram on the αε plane.

Reference
[1] Leggett A J Chakravarty S Dorsey A T Fisher M P A Garg A Zwerger W 1987 Rev. Mod. Phys. 59 1
[2] Weiss U 1993 Quantum Dissipative Systems Singapore World Scientific
[3] Shnirman A Makhlin Y Schon G 2002 Phys. Scr. T102 147
[4] Tong N H Vojta M 2006 Phys. Rev. Lett. 97 016802
[5] Rosenberg D Nalbach P Osheroff D D 2003 Phys. Rev. Lett. 90 195501
[6] Si Q Rabello S Ingersent K Smith J L 2001 Nature 413 804
[7] Seoanez C Guinea F CastroNeto A H 2007 Europhys. Lett. 78 60002
[8] Mühlbacher L Egger R 2004 Chem. Phys. 296 193
[9] Borri P Langbein W Schneider S Woggon U Sellin R L Ouyang D Bimberg D 2001 Phys. Rev. Lett. 87 157401
[10] Kehrein S Mielke A 1996 Phys. Lett. A 219 313
[11] Bulla R Tong N H Vojta M 2003 Phys. Rev. Lett. 91 170601
[12] Bulla R Lee H J Tong N H Vojta M 2005 Phys. Rev. B 71 045122
[13] Guo C Weichselbaum A vonDelft J Vojta M 2012 Phys. Rev. Lett. 108 160401
[14] Winter A Rieger H Vojta M Bulla R 2009 Phys. Rev. Lett. 102 030601
[15] Alvermann A Fehske H 2009 Phys. Rev. Lett. 102 150601
[16] Chin A Turlakov M 2006 Phys. Rev. B 73 075311
[17] Anders F B Bulla R Vojta M 2007 Phys. Rev. Lett. 98 210402
[18] Z G Zheng H 2007 Phys. Rev. B 75 054302
[19] Zheng H Z G 2013 J. Chem. Phys. 138 174117
[20] Florens S Freyn A Venturelli D Narayanan R 2011 Phys. Rev. B 84 155110
[21] Le Hur K Doucet-Beauprè P Hofstetter W 2007 Phys. Rev. Lett. 99 126801
[22] Yao Y Duan L Z Wu C Q Zhao Y 2013 Phys. Rev. E 88 023303
[23] Liu H B An J H Chen C Tong Q J Luo H G Oh C H 2013 Phys. Rev. A 87 052139
[24] Kast D Ankerhold J 2013 Phys Rev Lett 110 010402
[25] Nalbach P Torwart M 2012 Phys. Rev. B 81 054308
[26] Le Hur K 2012 Phys. Rev. B 85 140506
[27] Goldstein M Devoret M H Houzet M Glazman L I 2013 Phys. Rev. Lett. 110 017002
[28] Egger D J Wilhelm F K 2013 Phys. Rev. Lett. 111 163601
[29] Bulla R Costi T A Pruschke T 2008 Rev. Mod. Phys. 80 395
[30] Wilson K G 1975 Rev. Mod. Phys. 47 773
[31] Hou Y H Tong T H 2010 Eur. Phys. J. B 78 127
[32] Tong N H Hou Y H 2012 Phys. Rev. B 85 144425
[33] Bulla R Costi T A Vollhardt D 2001 Phys. Rev. B 64 045103
[34] T. Liu, Z.X. Cao, 2013 arXiv: 1312.4606 [quant-ph]
[35] Sassetti M Weiss U 1990 Phys. Rev. Lett. 65 2262
[36] Egger R Grabert H Weiss U 1997 Phys. Rev. E 55 R3809
[37] Volker K 1998 Phys. Rev. B 58 1862
[38] Costi T A Kieffer C 1996 Phys. Rev. Lett. 76 1683
[39] Keil M Schoeller H 2001 Phys. Rev. B 63 180302
[40] Florens S Venturelli D Narayanan R 2010 Lect. Notes Phys. 802 145
[41] Zhao C J Z G Zheng H 2011 Phys. Rev. E 80 011114
[42] Gan C Zheng H 2009 Phys. Rev. E 80 041106
[43] Vojta M Tong N H Bulla R 2005 Phys. Rev. Lett. 94 070604
[44] Vojta M Tong N H Bulla R 2009 Phys. Rev. Lett. 102 249904(E)