Abstract
We investigate the formation of magnetic Bose polaron, an impurity atom dressed by spin-wave excitations, in a one-dimensional spinor Bose gas. Within an effective potential model, the impurity is strongly confined by the host excitations which can even overcome the impurity-medium repulsion leading to a self-localized quasi-particle state. The phase diagram of the attractive and self-bound repulsive magnetic polaron, repulsive non-magnetic (Fröhlich-type) polaron and impurity-medium phase-separation regimes is explored with respect to the Rabi-coupling between the spin components, spin–spin interactions and impurity-medium coupling. The residue of such magnetic polarons decreases substantially in both strong attractive and repulsive branches with strong impurity-spin interactions, illustrating significant dressing of the impurity. The impurity can be used to probe and maneuver the spin polarization of the magnetic medium while suppressing ferromagnetic spin–spin correlations. It is shown that mean-field theory fails as the spinor gas approaches immiscibility since the generated spin-wave excitations are prominent. Our findings illustrate that impurities can be utilized to generate controllable spin–spin correlations and magnetic polaron states which can be realized with current cold atom setups.
Export citation and abstract BibTeX RIS
Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.
1. Introduction
Numerous aspects of quasiparticle physics, first introduced by Pekkar and Landau [1, 2], have been explored in a multitude of ultracold atom experiments [3–6]. Their implications range from condensed matter systems [7], e.g. organic semiconductors [8], to chemistry [9] and biophysics [10]. A widely studied quasiparticle in the cold-atom realm is the polaron [11, 12]; an impurity dressed by the elementary excitations of the bath with distinct bosonic and fermionic quantum statistics [5, 13–15]. Fascinating static features of these states include their effective mass [16, 17], induced interactions [18, 19] and formation of bound states [20–22]. Typical nonequilibrium phenomena associated with these structures revealed, among others, their induced correlations [23, 24], dynamical decay and relaxation [25–27], as well as transport [28–30].
In nearly all studies of polaron physics, the bath in which the impurity is immersed in is structureless. For the case of a bosonic host, which we are interested in, the amplitude of the phononic excitations triggered by the impurity are suppressed [24, 31], rendering the observation of the polaron cloud challenging. This complication can be alleviated by immersing the impurities into a spinor host where they act as the local perturbers modifying the underlying spin-order. These distortions of the magnetic environment give rise to spin-wave excitations [32] leading to the formation of magnetic Bose polaron. This quasiparticle is arguably far less explored than its fermionic counterpart [33–36], and non-magnetic polarons. A proposal to detect the polaron cloud, composed of many-body bound states in the strong-coupling regime, using interferometric techniques in a ferromagnetic gas was made [37]. Also, it was argued that a spinor bath can lead to polaronic subdiffusive behavior [38].
A spin-carrying medium features the premise of controlling the spin-order and stability of magnetic configurations by deploying an impurity as a probe. This point of view, we expect to provide a fresh impetus for studying impurity physics and novel many-body phenomena that would otherwise be challenging to detect. Of particular interest is the stability properties of ferromagnetic domains constituting a longstanding open problem in condensed matter physics [39]. Another promising direction is the understanding of the time-evolution and interactions of magnetic polarons, which is an active topic in spinor fermionic systems [35, 36]. They are thought to provide insights into the nature of high-temperature superconductivity [40].
An intriguing question is whether a crossover from a non-magnetic to a magnetic Bose polaron can be realized. Indeed, since the magnetic character of the medium can be in principle controlled via external radiofrequency fields, it might be possible to suppress spin-excitations by appropriately tuning the impurity-medium spin-dependent interactions, in relation to the imposed Rabi-coupling of the spin states. In fact, it has been demonstrated that impurity–impurity induced interactions can be tuned via the Rabi-coupling [41].
To shed light on the impact of an impurity subsystem in a magnetic environment and unravel the magnetic polaron characteristics, we deploy an impurity immersed in a one-dimensional (1D) harmonically trapped spin-1/2 Bose gas. We use the variational nonperturbative multi-layer multi-configuration time-dependent Hartree method for atomic mixtures (ML-MCTDHX) [42] to calculate the spin-order properties of an impurity immersed in a Bose gas. This method treats the impurity-gas system beyond the usual Bogoliubov treatment and thus the Fröhlich model [43], as well as accounts for boson–boson correlations beyond the Lee–Low–Pines approach [44].
We map out the 1D magnetic polaron phase diagram with respect to the impurity-medium coupling, the spin–spin interactions and the Rabi-coupling of the host spin constituents. In particular, attractive and self-bound repulsive magnetic polaron states are identified. The self-bound character of the latter stems from the pronounced magnetic excitations of the host binding the impurity within the gas despite the strong density–density repulsion. Additionally, the intervals where either the non-magnetic polaron takes place or the quasiparticle picture is no longer valid (due to medium-impurity phase-separation) are delineated. The latter two phases are known to also occur in the genuine three-component system [45, 46]. However, here we demonstrate that it is the presence of Rabi-coupling that proliferates the development of magnetization phenomena and consequently the magnetic polaron [37].
The standing spin-wave excitations dressing the impurity are captured by monitoring the spin-flips of each component of the medium. The magnetic character of the latter is inferred by the suppression of ferromagnetic spin–spin correlations. The latter, being susceptible to parameter variations, are found to be substantial near the miscibility–immiscibility transition of the host spin components. These manifest at specific interaction-dependent Rabi-couplings and importantly are tunable via the impurity-medium interaction. As such the impurity serves as a knob for manipulating the spin-order. The magnetic polaron residue decreases with increasing impurity-medium interactions illustrating increased magnetic dressing. Importantly, the residue suppression for large host particle numbers hints towards the susceptibility of the magnetic polaron to Anderson orthogonality catastrophe [47, 48]. The energy of the magnetic polaron is negative (positive) for attractive (repulsive) impurity-medium coupling and it increases for larger Rabi-coupling.
This work is structured as follows. Section 2.1 introduces the magnetic polaron setting and its intrinsic spin-symmetries while elucidating the role of the impurity to probe spin-wave background excitations. The phase diagram of the magnetic polaron is discussed in section 3 while its properties are analyzed in detail in section 4 including the associated spin–spin correlations and transfer processes. A summary of our findings together with future perspectives are provided in section 5. Appendix
2. Magnetic polaron setup
To study the magnetic Bose polaron, we use a three-component highly particle imbalanced bosonic setting. Specifically, a structureless impurity (I) atom is immersed in a spin-1/2 bosonic medium having N = N↑ + N↓ = 100 particles. The multicomponent setup is mass-balanced (unless stated otherwise), namely mI = mB , and all components are confined in the same 1D harmonic trap, ωB = ωI = ω. A corresponding experimental realization is possible by emulating the spin degrees-of-freedom of the medium e.g. via the Rabi-coupled hyperfine states |↑⟩ ≡ |F = 1, mF = −1⟩ and |↓⟩ ≡ |F = 2, mF = 1⟩ of 87Rb [49, 50]. The impurity atoms might refer to either the |F = 1, mF = 1⟩ state of 87Rb or the isotope 85Rb.
2.1. Impurity in a magnetic environment
The underlying many-body Hamiltonian
with the non-interacting parts of the spinor gas where α = ↑, ↓ and the impurity . The field operators act on the spin-a component of the Bose medium and on the impurity atom. Here, we consider ultracold temperatures ensuring the relevance of solely the s-wave interactions that are essentially described by a contact potential [51]. These are characterized by the spin–spin medium effective couplings gαα' with α, α' = ↑, ↓ and the impurity-spin interaction strengths gIa . In this sense, the contact spin–spin interaction term is , while the impurity-spin-a interactions are given by .
These effective coupling constants are related to the three-dimensional s-wave scattering length, (with σ, σ' = ↑, ↓, I being the species index) and the transversal confinement ω⊥ [51]. Consequently, the interaction coefficients can be experimentally adjusted either via Feshbach resonances [52, 53] through or utilizing confinement-induced resonances [51] by tuning ω⊥.
The spinor contribution facilitating magnetic processes reads
The Rabi-coupling ΩR favors the spin-superposition for the medium atoms and therefore admixes different numbers of spin-↑ and spin-↓ atoms in the ground state. The detuning δ = ν − ν0 provides the frequency difference of the Rabi-coupling laser from the corresponding non-interacting (g↑↑ = g↓↓ = g↑↓ = 0) resonance transition frequency (ν0) among the |↑⟩ and |↓⟩ states. Here, we consider δ = 0, corresponding to a resonant driving in the non-interacting case.
2.2. Spin-symmetries
The total spin operator is , where σ is the Pauli vector. Accordingly, the total spin magnitude operator is expressed as
where the so-called spin ladder operators . Importantly, the Hamiltonian of equation (1) preserves the SU(2) symmetry, i.e. only in the case of gI↑ = gI↓ and g↑↑ = g↓↓ = g↑↓. Otherwise, the eigenvalues of are not good quantum numbers. Also, in the presence of Rabi-coupling, ΩR ≠ 0, Sz is not conserved, allowing the system to become spin-imbalance, i.e. attain a superposition of |↑⟩ and |↓⟩. 8 Here, we focus on g↑↑ = g↓↓ ≡ g and therefore in the absence of the impurity, the system respects the symmetry implying invariance under spin-inversion along the z-spin axis. However, since the impurity can exhibit different interactions with each of the spin-components, namely gI↑ ≠ gI↓, breaking it can thus modify the magnetization of the system because Sz is broken.
2.3. Rescaling and suggested experimental procedure
In the following, the many-body Hamiltonian described by equation (1) is expressed in terms of ℏω. Then, the length, time, Rabi-coupling and interaction strengths are provided with respect to , ω, and respectively. Experimentally our 1D multicomponent setup is realizable, for instance, by using two hyperfine states of 87Rb (see above) with J m and a longitudinal (transversal) trap frequency ω = 2π × 100 Hz (ω⊥ ≈ 2π × 5.1 kHz). The respective temperature effects are suppressed as long as the condition μK is fulfilled [54]. In the latter expression kB is the Boltzmann constant, denotes the transversal confinement length scale and T refers to the temperature of the spinor Bose gas.
In an experiment, our results can be probed by coupling the spin components of the medium via a radiofrequency field [55]. The corresponding implementation involves an adiabatic ramp (within a time-interval ) of ΩR. The rf field creates a superposition spin state |↑⟩ and |↓⟩ that crucially depends on gαα' and gIα .
2.4. Phases of the spin-1/2 Bose gas in the absence of the impurity
In the absence of impurity, the considered setting reduces to a binary Rabi-coupled Bose–Einstein condensate (BEC) [49, 55–57]. The fundamental building block of such a Rabi-coupled mixture is the case of ΩR = 0, where it boils down to a two-component mixture with particle number conservation in each component. Then, it is well-known [58] that the system enters the phase-separated (miscible) state when (in the opposite case). Switching on the Rabi-coupling breaks the Sz
spin symmetry of the gas meaning that spin-transfer between the components is allowed. As a general rule the effect of the Rabi-coupling reduces the degree of immiscibility (or spin segregation) even when the spin interactions lie deeply inside the immiscible region. Specifically, for a larger ΩR the gas features a second-order phase transition entering the miscible state and spin-demixing is prohibited [49, 56, 57], see more details in appendix
2.5. Probing the spin order with the impurity
The spinor bosonic medium is initially in the ground state configuration, characterized by specific spin–spin interactions g↑↑ = g↓↓ ≡ g and g↑↓ and Rabi-coupling ΩR. In the absence of the impurity, i.e. gI↑ = gI↓ = 0 or equivalently , the spins are aligned and the associated spin configuration is almost 9 identical to a fully polarized state along the x-axis, see figure 1. It reads with i = 1, 2, ..., N. Then, for gI↑ ≠ gI↓ i.e. when the impurity-medium interaction is spin-dependent, the initial polarization vector , with ex the unit vector, rotates away from the x-spin axis.
Importantly, the localization of the impurity determined by its spatial width acts as a local perturber distorting the polarization of the medium. The local spin vector , for x ≠ x', is strongly influenced by the degree of localization of the impurity, see figure 1, in turn dressing the impurity with the spin fluctuations. This process gives rise to the magnetic polaron. An adequate measure to identify the degree of local perturbation in the bath is the spin–spin correlation function (see equation (13) below). For cases that the spin-order of the host remains intact by the presence of the impurity, the emergent quasiparticle will be referred to as a non-magnetic polaron.
To address the magnetic polaron properties, we employ the variational ML-MCTDHX approach which has been extensively used to study impurity dynamics and spectroscopy [23, 24, 26, 27, 45]. ML-MCTDHX is based on the expansion of the many-body wave function in terms of a time-dependent and variationally optimized basis set, see appendix
3. Phase diagram of the magnetic polaron
The phase diagram in figure 2 is calculated with the variational machinery of ML-MCTDHX. The interpretation of the different phase properties is greatly elucidated with an effective potential (see also appendix
The density of the α : {↑, ↓} spin components is ρα (x), see also equation (6). Typical effective potential approaches [24, 31] incorporate the density ρα (x) in the absence of impurity (gI↑ = gI↓ = 0) thereby neglecting all impurity-bath correlations. Here in order to account for the impurity backaction on the medium 10 , we construct an effective potential, , where ρα (x) corresponds now to the spin density obtained from the variational ML-MCTDHX approach. This comparison reveals the alteration of the quasiparticle phase diagram due to the above-mentioned correlations.
Download figure:
Standard image High-resolution imageFor simplicity, below, we focus on the case where gI↑ is finite, but gI↓ = 0. Initially we aim to analyze the competition between the Rabi-coupling ΩR and the impurity-spin-↑ interaction as shown in figure 2(a). We are particularly interested in the case , where spin-excitations dominate (see also appendix
To elucidate the main features of the different spin-orders that can emergent in the host, we first briefly analyze its phases in the absence of the impurity. In the Gross–Pitaevskii mean-field description for the binary Rabi-coupled bosonic bath [56], and within the Thomas–Fermi (TF) approximation (see appendix
below (above) which the host components are immiscible (miscible) for gI↑ = gI↓ = 0. Notice that n(0) = ρ↑(0) + ρ↓(0). Importantly, for this choice of interaction parameters, i.e. g↑↓ > g, the host becomes fully-polarized along the z-spin direction for ΩR = 0 [59–61]. Hence, the third ΩR regime appears which is characterized by negligible population of one BEC component. To estimate the extent of this regime, we define below which the minority medium spin-component is populated by less than one atom (see also appendix
Next we turn into the coupled impurity-spin-1/2 gas setting. Let us first consider the case where due to the strong effective magnetic field associated with ΩR, see equation (2), the host becomes fully-polarized along the x spin-axis. As such the BEC components are miscible implying that spin-excitations are rather inert, as discussed in reference [56]. Therefore, only very weak impurity-medium correlations can be induced. Then the medium can be viewed as an effective single-component BEC with a renormalized coupling (see also appendix
As , the magnetic degrees-of-freedom of the host gas become relevant. Here additional features appear due to spin-excitations. In particular, in the vicinity of the transition point between the repulsive Bose polaron and the phase-separated states at gI↑ ≳ g + g↑↓, another interspecies interaction regime appears. Here, the induced magnetic excitations modify the effective confinement of the impurity.
This effect is readily observed in figure 2(b4), where possesses a third well located at x = 0, in contrast to the double-well structure of Veff(x). This demonstrates the emergence of a magnetic Bose polaron state that is self-localized in the sense that the impurity is confined within the magnetic excitations it induces to its host. As , the magnetic polaron is stable for larger gI↑, see figure 2(a). Due to the interaction energy cost associated with the involved magnetic dressing cloud, this polaron is the ground state of the system 11 only for gI↑ < gmag c, see the dashed line in figure 2(a). For gI↓ < 0, the impurity also disturbs the polarization of its host giving rise to magnetic excitations inducing an additional effective attractive force, captured within , which lies beyond the Veff(x) picture, see figure 2(b1).
For , the host exhibits a phase separated character around x ≈ 0, leading to an attractive Veff(x) confining the impurity within the BEC independently of gI↑. Furthermore, the population of the spin-↑ component, which interacts with the impurity, is amplified (diminished) for gI↑ < 0 (gI↑ > 0), which aids the development of impurity-medium correlations. The pronounced magnetic excitations of the host, captured by , modify the effective confinement of the impurity Veff(x) and thus the polaron possesses a magnetic character. The latter ceases to exist for since the population of the minority spin-component becomes negligible. In particular, for repulsive gI↑ > 0, it is energetically preferable for the host atoms to occupy the non-interacting with the impurity spin-↓ component, thus preventing quasi-particle formation. However, gI↓ < 0, favors the spin-↑ configuration and as a consequence an attractive Bose polaron forms, akin to the single-component case [24, 26, 27].
Finally, we remark that for g↑↓ < g, no phase-separated regime for the host is encountered, see appendix
4. Impurity immersed in an immiscible spinor Bose gas
Since the spin-demixing processes in the case of a spin-1/2 Bose gas are enhanced for immiscible spin–spin interactions , our main focus is placed in this interaction regime for providing a clean manifestation of the magnetic polaron. As we argue in appendix
4.1. Magnetic polaron configurations and magnetization
To elucidate the impact of the impurity on the spatial distribution of the medium's spin components we invoke the one-body density of each component
Here, is the σ = ↑, ↓, I-component bosonic field operator acting at position x and |Ψ⟩ denotes the many-body ground state of the three-component system. This particle density can be observed by single-shot averaging [63]. Notice that , see also appendix
Download figure:
Standard image High-resolution imageFinite repulsive impurity-spin-↑ interactions lead to a depletion of ρ↑(x) around the trap center, see figure 3(a). Simultaneously, the impurity features a progressive localization tendency for larger gI↑ (figure 3(b)) thus forming a quasiparticle. This is the first key difference to the non-magnetic repulsive Bose polaron [24, 31] which is known to delocalize for increasingly repulsive impurity-host interactions causing its decay. Turning to gI↑ < 0, a behavior similar to the attractive non-magnetic Bose polaron [24] is detected. Here, the spin-↑ component shows a sizable density peak at the location of the impurity (figure 3(a)), with the concomitant localization of the impurity (figure 3(b)). However, the most striking distinction of the quasiparticle realized herein from the non-magnetic polaron stems from the response of the spin-↓ state. Indeed, ρ↓(x) is complementary to ρ↑(x), with the former being accumulated (depleted) in the spatial extent of the impurity for repulsive (attractive) gI↑ (not shown). This behavior of ρ↓(x) is mediated by the immiscible g↑↓ interactions which lead to an effective attraction, (repulsion, ), among the spin-↓ and the impurity for gI↑ > 0 (gI↑ < 0), see also figures 4(a5) and (c5) and the related discussion in section 4.2. In general, these effective interactions are mediated by the magnonic excitations due to the finite interaction difference gI↑ − gI↓. Notably, our observations necessitate the construction of effective Hamiltonians similar to [64] in order to appreciate the strength of induced interactions in three-component settings.
Download figure:
Standard image High-resolution imageThe above-described behavior of ρ↑(x) and ρ↓(x) hints towards the presence of magnetic processes in the system. These processes can be analyzed by studying the local magnetization response, provided in figure 3(c). The emergence of local magnetization in the host is related to a spatially varying effective magnetic field, Δ(x) ∼ (gI↑ − gI↓)ρI (x), due to the impurity-spin-↑ interactions (see also figure 1) which in turn gives rise to standing spin-waves. Without the impurity (gI↑ = 0), it holds that , due to the polarization of the host in the spin-x axis, while in the case of gI↑ > 0 (gI↓ < 0), we find . It is therefore the presence of the impurity that triggers a gI↑-dependent correlated standing spin-wave, which we analyze further in sections 4.3 and 4.4.
4.2. Two-body magnetic polaron correlations
A more concrete demonstration of the spatial distributions of the magnetic polaron is obtained by inspecting the underlying two-body configurations. For this reason, we determine the diagonal elements of the two-body reduced density matrix
This provides the probability to simultaneously detect a σ-component boson at position x1 and a σ' atom at x2. The spatially resolved two-body configurations for the immiscible interacting medium with ΩR/ω = 5 exhibit involved structures for both attractive (figures 4(a1)–(a5)) and repulsive (figures 4(b1)–(b5)) impurity-spin-↑ interaction strengths. Otherwise, for gI↑ = 0 they feature a symmetry under the exchange of x1 ↔ x2 and a localization around x1 = x2 = 0. This implies that the detection of the two atoms of the same or different components are largely independent, while the atoms are likely to reside around the trap center.
As explained above, an attractive impurity-medium coupling enforces the spin-↑ bosons to lie in the vicinity of the impurity. Consequently, is highly localized (figure 4(a1)) around the trap center demonstrating the portion of spin-↑ atoms bound to the impurity, see in particular (figure 4(a4)). The low density tails of infer the existence of spin-↑ bosons that remain unbound. On the other hand, exhibits localization in four disjoint spatial domains (figure 4(a2)) stemming from the phase-separation between the spin components which is evident by the elongated two-hump shape of (figure 4(a3)). As a consequence, the impurity and the spin-↓ are also phase-separated, despite being non-interacting gI↓ = 0 (see figure 3(b)). Therefore, shows a finite probability at two distinct domains characterized by |x↓| > |xI | (figure 4(a5)) which certifies the emergence of repulsive induced impurity-spin-↓ interactions caused by the attractive gI↑.
Turning to strong repulsive gI↑, an inversion in the roles of the spin-↑ and spin-↓ bosons takes place, compare e.g. figures 4(a2) and (b1). Indeed, in this case the spin-↑ atoms and the impurity tend to phase-separate (figure 4(b4)) due to their strong repulsion gI↑ ≫ g, while the spin-↓ particles are effectively attracted towards the impurity (figure 4(b5)), i.e. besides the fact that gI↓ = 0. Similar to the non-magnetic polaron case [24, 31], we observe that also here the quasiparticle character in the repulsive branch is less pronounced than in its attractive counterpart. Hence, the emergent patterns in figure 4(bi ) are less prominent than those depicted in figure 4(ai ) with i = 1, ..., 5.
4.3. Dressing cloud and waveform of the magnetic polaron
To reveal the imprint of the impurity into the spatial configuration of the spinor gas, we operate in the co-moving impurity frame. This is the natural frame of reference to measure the magnonic or phononic dressing cloud since it avoids effects stemming from the dispersion of the impurity within the confining potential. Such a procedure has been already successfully implemented in the experiment for monitoring the internal structure of magnetic Fermi polarons [65]. Focusing on the bosonic case described above and in order to extract the spatial distribution of the polaron dressing cloud, we consider the following two-body density ansatz
where , are the background 12 densities of the spin-α component and the impurity respectively. These do not directly contribute to the binding of the impurity to the host excitations, but account for the density inhomogeneity originating from the harmonic confinement. The function uα (xα − xI ) captures the impurity dressing by the atoms of its host. It is assumed to be solely a function of xα − xI , due to the short range character of the impurity-host interactions. As such, it results in significant variations of the BEC density only for |xα − xI | ∼ ξα , where ξα is the healing length of the spin-a component. For a homogeneous system and , where L is the length of the system, and thus reduces to the form expected within the Lee–Low–Pines mean-field description [16, 44, 66]. Therefore, we anticipate that equation (8) describes adequately the structure of only in the case that the local density approximation is justified, i.e. ξα ≪ Rout.
Provided that equation (8) is valid we can extract information regarding the dressing cloud of the impurity by employing the impurity-spin-α correlation function in the comoving frame xr = xα − xI , namely
Then is related to the existence of spin-wave excitation (magnon) dressing, captured by u↑(xr ) − u↓(xr ). Moreover, is associated to the presence of phonons. Note that for the case of a miscible spinor host (here achieved for ΩR = 5ω) it holds that for gI↑ = 0. Therefore, we expect that any deviation in the magnetization of the system for gI↑ ≠ 0 stems from the magnetic dressing cloud of the impurity. This implies that and consequently . In this way, captures the waveform of the standing spin-wave dressing cloud. We present for ΩR = 5ω with respect to gI↑ in figure 5(a). It features for gI↑ > 0 (gI↑ > 0) indicating the respective magnetization tendency. The shape of is found upon fitting to be well-described by
with Amag (w) denoting its amplitude (width). We expect that by operating in the Lee–Low–Pines framework one should be able to extract equation (10) as the solution of the magnetic polaron dressing cloud, an analysis that will be addressed in a future work.
Download figure:
Standard image High-resolution imageOn the other hand, it is more involved to extract the phononic dressing cloud as the non-vanishing contribution from the last term of equation (9) implies . To account for this correction we employ the following spin-independent pair-correlation ansatz
Here, w' refers to the width of the polaron and Aph is the amplitude of the phonon dressing cloud. In the above expression we have assumed a TF (Gaussian) background density for the BEC (impurity). With this choice the background two-body density, nTF(xr ), reads
where corresponds to the TF profile of the spin-α BEC component (see also appendix
Upon fitting the ansatz of equation (10) to and the one of equation (11) into we determine the width and amplitude of the magnetic polaron and the non-magnetic polaron respectively. Next, in order to discern the dressing cloud stemming from the magnons and the phonons we find the number of medium atoms lying in the waveforms u↑(xr ) ± u↓(xr ), see equations (10) and (11). These populations, and depicted in figures 5(b) and (c) feature an increasing tendency for a larger magnitude of impurity-spin-↑ interactions testifying quasiparticle formation. It can also be inferred that independently of gI↑, the magnonic excitations prevail over the phononic ones. This means that the emergent quasiparticle, being genuinely dressed by an admixture of magnons and phonons, possesses a dominant magnetic component. Particularly, the magnetic dressing cloud acquires a maximal value of about 30% of the medium atoms around , while the phonon branch is at most 3.5%.
For , the magnonic excitation branch is again pronounced as compared to the phononic one especially for gI↑ < 0, while both branches are suppressed for gI↑/g > 3.8 where phase-separation occurs. Naturally, if ΩR = 0 magnetic dressing is diminished for every gI↑, while the phononic one is finite for gI↑ < 0 since only in this case the impurity lies within the spin-↑ host. Regarding , we observe that Nmag is suppressed. In this scenario, the spin degrees-of-freedom are nearly frozen because of the large energy gap for exciting spin-waves [56]. Concluding, from the above we can deduce that the magnonic cloud should be easier experimentally detectable when compared to its phononic counterpart.
4.4. Broken spin-order and spin–spin correlations
To inspect the rise of spin-fluctuations in the medium and thus attest the emergence of the magnetic polaron we track the spin–spin correlation [68]
Here, is the total-spin operator of the system defined in equation (3). This correlation function probes the alignment among two spins and distinguishes ferromagnetic , antiferromagnetic and paramagnetic spin configurations 13 . It is showcased in figure 6(a) for immiscible spin–spin interactions and several Rabi-couplings as a function of the impurity-spin-↑ interaction strength. As it can be readily seen, for suppressed Rabi-coupling (ΩR = 0) the medium remains ferromagnetic (i.e. ) irrespectively of gI↑. Recall that the interacting eigenstate of the multicomponent setting corresponds to the one where all atoms are in the spin-↑ (↓) state for gI↑ > 0 (gI↑ < 0). In this scenario, a polaron dressed by the phononic excitations of the bosonic bath occurs as long as gI↑ < g + g↑↓. Otherwise, an impurity-medium phase-separation takes place evincing the polaron decay [24].
Download figure:
Standard image High-resolution imageIn sharp contrast, switching on the Rabi-coupling results generally to the suppression of the ferromagnetic order as evidenced by the reduction of for increasing (figure 6(a)). Particularly, spin correlations become enhanced within the critical magnetization region where spin-mixing is dominant (appendix
A similar to the above-described response of the spin-fluctuations takes place for a heavy impurity (mI ≫ mB ) or finite impurity spin-↓ couplings, see figure 6(b). Notice that for gI↓/g > 1.6, since then a phase-separation between the impurity and the spinor medium is favored as long as gI↑ > g + g↑↓ − gI↓.
4.5. Spin transfer processes
Having exemplified the role of spin-fluctuations caused exclusively by the impurity-medium interactions, we next aim to unravel the accompanied spin-demixing mechanisms. These refer to the migration of spin-↑ to spin-↓ particles and vice versa leading ultimately to interaction dependent spin configurations. The latter naturally provide further evidence of the presence of spin-wave excitations identified in figure 5(a). These are quantified herein by the portion of the spin-flipped atoms with respect to the gI↑ = 0 configuration. To capture the spin-demixing of the bosonic medium, due to gI↑ ≠ 0, for various Rabi-couplings ΩR, we monitor the fraction of bosons in each spin-α component
where N = N↑ + N↓. The respective fraction of spin-α atoms for several ΩR is depicted in figure 7(a). In the miscible regime, i.e. , it holds that , i.e. . For gI↑ > 0 (gI↑ < 0), ΔP decreases (increases). In the case of and gI↑ > g + g↑↓, due to phase-separation, see also the phase diagram of figure 2(a) and the effective potential depicted in figure 2(b3). For instance, at , the initial decreasing trend of for larger gI↑/g > 0 is followed by an abrupt saturation towards around gI↑/g = 1 + g↑↓/g = 2.6, due to the phase-separation between the impurity and its spinor environment in this region; see light and dark blue lines in figure 7(a). A similar abrupt behavior with respect to gI↑/g is also visible in other observables such as the residue (figure 8(a)) and the energy (figure 9). In the immiscible scenario, , even for gI↑ = 0 and it follows the same behavior as in the miscible regime with gI↑. For ΩR = 0, , because the spinor two-component host reduces to a single-component since (see also figure 2).
Download figure:
Standard image High-resolution imageDownload figure:
Standard image High-resolution imageDownload figure:
Standard image High-resolution imageThis behavior of in the miscible regime (here ΩR = 5ω) is not altered also for a finite value of gI↓ as long as gI↓ < g, see figure 7(b). Of course, due to gI↓ ≠ 0, the population balance scenario is achieved for gI↑ = gI↓. This becomes more pronounced for gI↓/g = 1, where we also observe that when gI↑ > g↑↓, see appendix
4.6. Magnetic polaron residue and energy
The quasiparticle residue [11] refers to the overlap between the non-interacting (gIα = 0) and the interacting (polaronic) state
with α = ↑, ↓. It can be experimentally tracked e.g. via radiofrequency spectroscopy [3, 5]. The behavior of with respect to gI↑ and distinct ΩR is presented in figure 8(a). Apparently, for ΩR = 0 the residue for gI↑ > 0 implying that the impurity is not dressed by its environment, thus corresponding to a free particle. This is in contrast to gI↑ < 0, where reduces for stronger attractions being a consequence of the attractive polaron dressed by the phononic excitations of the bosonic medium. Turning to larger ΩR, in the vicinity of , we find that reduces for increasing signifying a tendency towards a completely deformed interacting state. For repulsive gI↑ it is caused by the phase-separation, while for attractive interactions it stems from the impurity-medium bound state. This overall behavior of appears to be similar for a heavy impurity. For , the polaronic residue is dramatically different. Focusing on gI↑ < 0, it decreases for larger attractions but it is always larger than for ΩR/ω = 5. However, repulsive impurity-spin-↑ interactions result in a smooth decrease of for gI↑ < g + g↑↓. Otherwise a sharp reduction of takes place which is associated with the phase-separation among the impurity and the spin-↑ state. A similar response is observed for a heavy medium and ΩR/ω = 5.
We then compare our variational findings with the results from the mean-field approximation. Such a direct comparison is especially motivated by recent studies on the 1D Bose polarons whose ground state characteristics have been argued to be described (at least to some extent [69–71]) within a mean-field framework [67]. This issue has been carefully benchmarked, for instance, against quantum Monte Carlo techniques [21], the Lee–Low–Pines transformation [17, 44] as well as the ML-MCTDHX and the flow equation (IM-SRG) methods [69, 70].
The respective mean-field approximation for our system is described by the following system of three coupled GPEs
where . The chemical potential of the medium μB
is inherently related with the particle number
14
N = N↑ + N↓, see also appendix
Indeed, deviations from the mean-field become in particular noticeable for repulsive impurity-medium interactions and around , implying that in this region spatial correlations become non-negligible. The imprint of spatial correlations on is especially pronounced by considering stronger interparticle interactions, compare the many-body and the mean-field results in figure 8(b). In particular, the presence of correlations lead to smaller residue for repulsive gI↑ due to the superposition of the many-body wave function as compared to the mean-field value. For attractive gI↓, the situation is reversed because the impurity is less spatially localized in the mean-field case.
Since the mean-field solution is adequate for weak medium interactions, we have also studied the behavior of for substantially larger atom numbers. For this investigation, we systematically approach the thermodynamic limit 15 observing that , where |ϕσ ⟩ and |ϕσ0⟩, with σ ∈ {↑, ↓, I} are the mean-field order parameters for gI↑ ≠ 0 and gI↑ = 0 respectively. Thus, , for large NB and NI , where corresponds to the residue for N = 100. This exponentially decaying behavior of manifests the Anderson catastrophe of the magnetic polaron in the thermodynamic limit in 1D [11, 47]. It further supports the generalization of our results and their experimental detection in the large atom number limit as long as the impurities density is low to ensure that their interactions are negligible.
The energy of the quasiparticle is naturally determined by the energy difference among the interacting impurity-medium setting and the non-interacting one,
As it can be verified by inspecting figure 9, the polaron energy shows a continuously increasing (decreasing) trend for stronger repulsive (attractive) gI↑ as long as ΩR ≠ 0. Moreover, for and in the case of gI↑ > g + g↑↓, the energy saturates due to the impurity-spin-↑ phase-separation process (see also figure 3). The saturation value of the polaron energy caused by the impurity-medium phase-separation can also be predicted within Veff(x) (see equation (C4) in appendix
5. Conclusions and outlook
We have investigated the ground state properties of a structureless impurity embedded in a spin-1/2 Bose gas. The interaction of the impurity with one spin component is switched on, leading to spin-wave excitations of the host atoms and formation of magnetic Bose polarons. To evaluate beyond mean-field correlations in the magnetic polaron formation, we compare results obtained within the variational approach and the three-component GPEs. The spin-spatial correlations play the dominant role for weak boson–boson interactions, while spatial correlations become important for stronger interparticle couplings, invalidating the mean-field treatment. An effective potential for the impurity immersed in the bosonic bath is constructed to elucidate the polaron characteristics. Interestingly in the magnetic polaron regime, the impurity is confined within an effective attractive potential well originating from the host excitations it triggers. Particularly, in the repulsive branch, this emergent attraction stabilizes the quasi-particle against impurity-bath phase separation.
The phase diagram of the 1D magnetic polaron, as a function of impurity-medium, spin–spin interactions, and Rabi coupling among the spin components is calculated. The new phases include attractive and self-bound repulsive magnetic polaron configurations, and impurity-medium phase-separated regions where the quasiparticle decays. The transition from non-magnetic to magnetic polaron states strongly depends on the Rabi coupling. The spin-wave excitations emerge as a localized disturbance of the host magnetization, affecting the occupation of the individual spin components. This mechanism is tunable by varying the impurity-medium interactions while keeping all other system parameters fixed. In the absence of impurity, it is explicated that the binary magnetic gas can be efficiently driven from a miscible to a spin-segregated phase via Rabi tuning. It is in the vicinity of this phase transition where the magnetic properties of the polaron become more prominent.
We show that spin–spin correlations are affected with finite impurity-medium coupling, offering the opportunity to use impurity-spin interaction to control the local bath spin order. To gain further insights into the magnetic polaron states, we calculate the residue, showing a decreasing trend for large impurity-medium coupling. The presence of spatial correlations for strong boson–boson interactions results in a suppressed residue behavior. Moreover, by inspecting the polaron energy, it is shown to be negative for attractive impurity-medium couplings, indicating the formation of an impurity-medium bound state. The latter is less bound for larger Rabi coupling, while for repulsive interactions, it experiences an increasing tendency.
It would be intriguing to develop an analytical understanding of the impact of correlations leading to the renormalization of the effective potential experienced by the impurity in the magnetic polaron regime. Moreover, the impact of the impurity-spin interaction on the critical behavior and associated universal exponents across the miscible–immiscible phase transition in the large atom limit, is a fruitful pathway. In particular, this investigation can provide additional insights into the properties of the magnetic polaron appearing in the critical regime.
Another immediate prospect is to examine the quench dynamics of the magnetic polaron following a time-dependent ramp of the intensity of the radiofrequency field (Rabi coupling) in order to examine the possibility of spin domain formation. Along these lines, it is important to emulate radiofrequency or Ramsey spectroscopy in the spinor medium for unveiling the many-body properties of its collective excitations, such as spin-waves. Another interesting direction would be to study the induction of spin-order in the presence of a spin–orbit coupled impurity atom. The generalization of our findings to two-dimensions where long-range effective spin-order, exhibiting also anisotropic character, is a worthy further pursuit.
Acknowledgments
SIM and HRS acknowledge support from the NSF through a grant for ITAMP at Harvard University. GMK and PS have been funded by the Cluster of Excellence 'Advanced Imaging of Matter' of the Deutsche Forschungsgemeinschaft (DFG)-EXC 2056-Project ID 390715994. FG acknowledges funding by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany's Excellence Strategy—EXC-2111—390814868. This work was also financially supported within the PIER Hamburg-MIT/BOS Seed Projects with the Project ID: PHM-2019-06.
Data availability statement
The data generated and/or analysed during the current study are available from the corresponding author on reasonable request.
Appendix A.: Impurity in a miscible spinor Bose medium
Below, we showcase that the polaron exists also in the case of a miscible interacting spinor medium, however, its magnetic character depends crucially on ΩR. Here, we assume g = 0.5 and g↑↓/g = 0.9 while gI↓ = 0. The σ-component densities are provided in figures 10(a)–(c) for different impurity-spin-↑ interactions and in the case of ΩR = 0. For repulsive gI↑ we observe that in the region gI↑ < g↑↓ + g the impurity cloud widens for larger gI↑ while remaining within ρ↑(x). Entering gI↑ > g↑↓ + g the impurity moves to the edges of the TF cloud of the medium, x = ±Rout, splitting into two symmetrically placed density branches minimizing its overlap with ρ↑(x). However, ρ↓(x) slightly widens as a result of its effective attraction mediated by the repulsive gI↑, see section 4.1. The local magnetization in the vicinity of the impurity is negative in this repulsive interaction region manifesting the existence of a magnetic polaron (figure 10(d)). The magnetic quasiparticle nature is also supported by the occurrence of spin-fluctuations identified since as shown in figure 10(f). Notably, irrespectively of gI↑ for increasing ΩR implying that the magnetic character is diminished. This is expected as for increasing ΩR the host becomes progressively more strongly polarized along the spin-x axis and therefore the impurity can hardly induce spin-demixing. Consequently, the atoms are distributed in an almost equal fashion between the spin components, see figure 10(e). The latter feature a small population imbalance being more pronounced before the overlap among the impurity and the spin-↑ bosons becomes minimal and it is further suppressed for larger ΩR.
Download figure:
Standard image High-resolution imageTurning to attractive gI↑, a non-negligible portion of the spin-↑ atoms accumulate close to the impurity and a self-bound attractive magnetic polaron forms. Simultaneously, ρ↓(x) tends to separate with ρ↑(x) as a result of the repulsive induced interactions mediated by the attractive gI↑. Here, is positive in the location of the impurity (figure 10(d)) verifying the magnetic character of the polaron which can also be inferred by the fact that (figure 10(f)). Once again the reduction of is smaller for a larger ΩR indicating the suppression of the magnetic polaron character. The population transfer is enhanced compared to the repulsive case (figure 10(e)). This implies that the polarization is not adequate for characterizing spin–spin correlations and vice versa [68].
Appendix B.: Phase structure of the two-component magnetic Bose gas
Let us briefly outline the mean-field ground state characteristics of a binary magnetic bosonic gas, analyzed to some extent in reference [56]. A simplified description of the pseudospinor BEC is obtained by relying on the TF approximation in the three-component GPE description (equation (16)). Particularly, the kinetic energy of the medium is neglected leading to the following coupled set of equations for the medium
where ɛ0(x) ≡ (ɛ↑(x) + ɛ↓(x))/2, Δ(x) ≡ ɛ↑(x) − ɛ↓(x) and , with α ∈ {↑, ↓}. Here we are interested in the magnetic properties of the BEC in the absence of the impurity and therefore we have set ψI (x) = 0 in equation (16).
Formally any self-consistent solution of equation (B1) is an adequate solution of the GPE within the TF approximation. However, herein, we are interested in the ground state of the system and as a consequence we minimize the energy contribution stemming from yielding for all x. In this case, equation (B1) and Δ(x) = ɛ↑(x) − ɛ↓(x) define two algebraic equations for ɛα (x), α ∈ {↑, ↓} the solution(s) of which determine the ground state(s) of the pseudospinor BEC.
However, in order to identify the distinct phases of the Bose gas in a more transparent manner it is more convenient to work with quantities based on its wave function (and thus also its density). Indeed, the choice motivates us to express the wave function as
with n(x) = ρ↑(x) + ρ↓(x) denoting the total BEC density. Accordingly, the phase factor reads
Equation (B2) expresses the wave function of the system solely in terms of the functions n(x) and Δ(x). To render clear the physical interpretation of these functions, we note that the densities of the individual BEC components are
As such, Δ(x) signifies the degree of local spin-imbalance among the distinct components. In this sense Δ(x) ≠ 0 indicates spin-component immiscibility.
To derive the ground state solutions of the pseudospinor Bose gas we insert equations (B2), (B3) into equation (B1) resulting in the self-consistency equations
Notice here the appearance of the chemical potential, μB , which is fixed by demanding N = ∫dx n(x), yielding a third equation for obtaining the TF profile. For simplicity, below, we consider a two-component system with g↑↑ = g↓↓ = g. Therefore, the solutions of equation (B5) read
where Rin denotes the TF radius of the spin density imbalance, ρ↑(x) − ρ↓(x) = −Δ(x)/(g↑↓ − g), and Rout is the TF radius of each component. Also, the chemical potential μB refers to the solution of the following algebraic equation
The solutions provided in equation (B6) reveal that the pseudospinor BEC experiences a phase transition in terms of ΩR and g↑↓. For large Rabi couplings, i.e. , a unique solution exists (since Rin = 0 and the first branch does not contribute) characterized by a completely miscible phase with Δ(x) = 0 and therefore ρ↑(x) = ρ↓(x). In contrast, reducing the Rabi coupling such that the ground state is characterized by the coexistence of two separate spatial domains in terms of Rin and Rout. In particular, since in this case Rin > 0, there is an immiscible region close to the trap center, namely for |x| ⩽ Rin. Since the system is symmetric to spin inversions, namely it exhibits a symmetry (see discussion in section 2.2) owing to the fact that g = g↑↑ = g↓↓, there are two distinct solutions in this regime. These possess either Δ(x) < 0 or Δ(x) > 0, corresponding to a predominant occupation of the spin-↑ or the spin-↓ state at the trap center respectively, see equation (B4). Therefore, we conclude that in the immiscible case the ground state of the BEC is doubly degenerate. Furthermore, assuming a non-zero Rabi coupling, ΩR ≠ 0, we obtain Rin < Rout and therefore a miscible region always appears for Rin ⩽ |x| ⩽ Rout. In this spatial extent the densities of the components are perfectly overlapping, as in the case discussed above. Notice that this partially immiscible regime does not appear for g↑↓ < g, since the state of the Bose gas is miscible even for ΩR = 0. Indeed, in this case it is impossible to satisfy the immiscibility condition , because n(0) > 0 and |ΩR| > 0.
To gain deeper insight into the behavior of the pseudospinor BEC for varying ΩR and g↑↓ we present its underlying phase diagram in figure 11(a) for the parameters employed in the main text, namely g = 0.5 and N = 100. Apparently, the partially immiscible phase emerges for strong g↑↓ and weak ΩR. Furthermore, the variation of Rin and Rout for g↑↓ = 0.8 is illustrated in figure 11(b). It can be readily verified that Rin = Rout holds for ΩR = 0, while an increasing Rabi-coupling leads to a decrease of Rin, ultimately reaching Rin = 0 for . In contrast, Rout increases compensating for the constant particle number (N) in the BEC. The miscible–immiscible phase transition for larger ΩR is directly captured by the population of the individual component Nα = ∫dx ρα (x) depicted in figure 11(c). As it can be seen within the TF approximation and for ΩR = 0 the system is fully polarized along the z spin axis. For higher values of ΩR the particle imbalance decreases and eventually vanishes within the miscible regime.
Download figure:
Standard image High-resolution imageIt should be, however, stressed that the TF approximation is not quantitatively reliable in the immiscible regime since the derivative of the component wave functions, Ψα (x), is discontinuous at |x| = Rin. 16 Therefore, the kinetic energy of the BEC becomes infinite in the vicinity of these points. In order to correctly evaluate the BEC density profiles of the partially immiscible system one has to solve the full GPE equations of motion (equation (16)) and thus account for the impact of the kinetic energy term. Indeed, upon considering the GPE model, Nα is altered in the vicinity of the transition region from the TF results, see figure 11(c). Nevertheless, the TF approximation is found to provide an adequate qualitative description of the phase of the Bose gas. Below, we shall employ this approximation to elucidate the phase diagram of the impurity immersed in a pseudospinor BEC.
Appendix C.: Effective potential and its range of applicability
Previous studies focusing on a spinless BEC host [24, 26, 27, 31] argued that important insights regarding the equilibrium and the dynamical Bose polaron properties are obtained by merely considering the Bose gas acting as a material barrier to the impurity. This is the concept of the so-called effective potential which of course neglects correlations among the impurity and the medium. Its success can be traced back to the separation of energy scales between density–density interactions giving rise to the effective potential and the corresponding bath-phonon interactions predicted within the Fröhlich model [72, 73]. Indeed, the energy scale of density–density interactions is Eden–den ∼ gBI n0, with gBI being the impurity-medium effective coupling strength and n0 denoting the BEC density. In contrast, the impurity-phonon coupling within the Fröhlich model is [72], where ξ is the BEC healing length. Thus, the energy scale defined by the density–density interactions is larger, since for a BEC the healing length should be larger than the corresponding interparticle distance [74]. Consequently, the impact of the effective-potential is more pronounced than the phononic dressing.
The generalized form of the effective potential for the case of a pseudospinor BEC reads
where ρα (x), with α ∈ {↑, ↓}, is the density of the corresponding BEC component for gI↑ = gI↓ = 0. To establish the limitations of the effective potential approach, Veff(x) is also compared to the improved effective potential . The latter possesses the same form as Veff(x) but ρα (x) corresponds to the BEC density calculated for the proper values of gI↑, gI↓ within ML-MCTDHX. Therefore, incorporates effects stemming from the impurity-medium correlations allowing us to characterize their impact in the polaronic state and identify effects beyond the effective potential framework. This inclusion is important due to the pseudospin degree-of-freedom of the BEC. As shown in reference [56] except for the phononic excitations (which retain the same structure as in the spinless scenario within the miscible regime, ) a new branch of spin-excitations emerges referring to out-of-phase density fluctuations among the BEC components. Here we are not interested in a detailed study of the coupling mechanism between spin-excitations and the impurity. However, such contributions lead to pronounced impurity-medium correlations beyond Veff(x). These can be directly captured by comparing Veff(x) with .
Below, we derive the explicit form of Veff(x) (equation (C1)) within the TF approximation utilizing the results of appendix B, for both miscible and immiscible BEC components. This exploration will allow us to interpret the phase diagram of the different polaronic excitations emanating for repulsive impurity-medium interactions but also unravel the limitations of this effective approach.
C1. Miscible regime
Referring to miscible spinor components (Rin = 0), according to equation (B6), the effective potential reads
here the factor Dmisc determines the shape of the effective potential and reads
Therefore, Veff(x) possesses a harmonic oscillator shape for Dmisc > 0, or equivalently deforming into a double-well in the opposite case, Dmisc < 0, with minima located at xmin = ±Rout. As such, in line with the results of references [24, 27] for a single-component host, we expect that the impurity and the medium bath become immiscible in the case of gI↑ + gI↓ > gBIcrit. This ultimately results in the instability of the repulsive Bose polaron being inherently related to its decaying character in this interspecies interaction regime [24, 31]. This provides the first threshold for the stability of the Bose polaron indicated in figure 12(a), see in particular the edge of the blue region for g↑↓ = gI↑ − g. This is equivalent to the edge of the repulsive Bose polaron regime for gI↑ = g + g↑↓ appearing in figure 2(a).
Download figure:
Standard image High-resolution imageIt should be emphasized that within Veff(x) we completely neglect backaction effects of the impurity to the BEC, which in the case of a pseudo-spinor Rabi-coupled BEC can be important for determining the stability of the Bose polaron. Nevertheless, by comparing Veff(x) with the corrected potential incorporating the effect of impurity-medium correlations we do not observe significant deviations among the two approaches as long as, the Rabi-coupling is strong enough, see figure 2(b2) for the repulsive polaron and figure 2(b3) for the immiscible impurity-medium case. Similar findings are obtained for g↑↓≪ g (not shown here for brevity). The above can be interpreted in view of reference [56], where it was demonstrated that the spin-excitations possess a sizable gap when the system is deep into the miscible regime of the pseudospinor BEC. Therefore, within the latter regime the generation of spin-wave excitations into the host is prohibited. Recall that also the phonon-impurity coupling is much weaker than the Veff(x) contribution, thus suppressing possible density excitations of the medium. As such, both excitations pathways are essentially frozen within this miscible BEC regime and consequently the impurity-medium correlations are expected to be insignificant. This explains the fact that as shown in figures 2(b2) and (b3).
As the miscibility–immiscibility threshold is approached, i.e. for or , the above-mentioned energy gap among the spin and the phononic excitations closes [56] allowing for the coupling of the impurity state with the spin-fluctuations of its host. As a consequence we expect pronounced impurity-medium correlations in this regime. Indeed, by incorporating this correction into a new effective potential, , we observe the emergence of an additional well around x = 0 which is responsible for binding the impurity into the Bose gas provided that , see figure 2(b4). This behavior is reminiscent of the so-called Pekkar polaron according to which the density excitations of the host are maximized in the vicinity of the impurity leading to its spatial localization [2, 75]. Importantly though, in our case the excitations dressing the impurity possess a magnetic character (see section 4.4) and therefore we refer to the structure emerging in this regime as the Pekkar-magnetic polaron.
C2. Immiscible spin interactions
For a medium with immiscible spin components one has to consider a larger number of cases since depending on the sign of gI↑ − gI↓ distinct medium configurations are favored. Indeed, if gI↑ > gI↓ (gI↑ < gI↓) it is preferable that n↑ > n↓, Δ(x) < 0 (n↑ < n↓, Δ(x) > 0). This leads to a discontinuity in nα at gI↑ = gI↓. Then the effective potential in the TF approximation becomes
To further elucidate the form of Veff(x) we perform an expansion with respect to x around |x| ⩽ Rin. Within the harmonic approximation (i.e. dropping terms ∝xn for n > 2) we obtain
Consequently, whether Veff(x) exhibits a double- or single-well structure depends on the sign of
Indeed, Dim > 0 designates a single-well potential around x = 0, while Dim < 0 implies a double-well with minima at x = ±Rout. A first simplification occurs in the case of gI↓ = 0 (due to the symmetry we can equivalently assume gI↑ = 0), where Dim > 0 for all gI↑. This means that in the immiscible case and for gI↓ = 0, the impurity can not escape from the BEC within the TF approximation and the polaron exists for every gI↑.
By comparing, Veff(x) with within the immiscible regime, we observe a qualitative agreement between these two approaches see figure 2(b5). The notable differences lie in the modification of the effective potential around x = 0 owing to the magnon dressing. This leads to a stronger binding of the impurity in the BEC when impurity-medium correlations are incorporated. As such, the Bose polaron in this parameter range possesses a similar structure as the magnetic one identified in the miscible regime, despite that the former does not show a self-bound character. For this reason we do not differentiate among these two regimes in the phase diagram of figures 2(a) and (b) and refer to the emerging structures as magnetic Bose polarons. An additional alteration of occurs near x = ±Rin, originating from the non-negligible kinetic energy of the bosons in this spatial region which is neglected within the TF approximation. Further, for gI↑ = 0 (or gI↓ = 0), we note that deep in the immiscible regime, i.e. for (or equivalently for ) the BEC tends to be fully polarized. Then, all atoms occupy the spin-state that is non-interacting with the impurity and therefore in this case no polaron exists. Of course, as the system approaches this regime it exhibits a crossover character and consequently no phase-boundary emerges. In order to estimate the parameter region where the behavior of the system changes we define, possessing the property that for and fixed ΩR less than a single BEC atom occupies the component that is interacting with the impurity, see figure 12(a). Similarly, we have defined , see figure 2(a), such that for and fixed g↑↓ the component interacting with the impurity is occupied by less than one atom.
Let us now briefly comment on the case of gI↑ ≠ 0 ≠ gI↓, by examining the expected regimes where a polaron appears within the TF approximation (i.e. Dim > 0). The behavior of the system is the simplest for , where the spin components are strongly imbalanced and almost only one of them is occupied (see also figure 11). It naturally follows from equation (C6) that the polaron exists for . These results are in agreement to the ones regarding an impurity immersed in a spinless BEC [24, 27, 31]. On the contrary, in the region , i.e. close to the miscibility–immiscibility threshold, Dimm < 0 signifying that a potential minimum at x = 0 always exists independently of the impurity-medium interaction strength. Finally, in the intermediate case, , there is a threshold
below which the Bose polaron exists. Here, the function λ(ΩR) features an increasing behavior for , and exhibits the extrema λ(ΩR = 0) = 0 and . The precise values of λ(ΩR) in the above-mentioned domain, depend solely on the parameters characterizing the medium, namely g, g↑↓ and N. For illustration, figure 12(b) provides λ(ΩR), alongside with an example stability phase diagram in the inset, for the host parameters employed within this work.
Concluding, Veff(x) provides insight into the existence of polaronic states, but not whether they correspond to the system's ground state. Indeed, if the impurity remains within the BEC its energy increases without bound for a larger gIα . In contrast, the phase-separated states possess an energy that saturates for strongly repulsive impurity-medium interaction strengths since the interspecies spatial overlap is small. Therefore, it is expected that the impurity energy in these cases will saturate towards . This implies that even in the cases that the Bose polaron persists for large interactions, the phase-separated states might be the overall ground states of the system and therefore explicit investigations are required to determine the energy of both branches. Here a hint towards the stability of the polaronic states relies on the fact that the state of the BEC is highly perturbed for the cases of the magnetic polaron and almost completely intact for a phase-separated impurity-medium configuration. As a consequence, the corresponding coupling among the polaronic and phase-separated states that would render the former configuration unstable is small. This can be understood from the fact that the state overlap typically scales as , where ρα (x), are the original and perturbed single-particle densities respectively.
Appendix D.: Variational treatment of the magnetic polaron
In order to capture the correlation properties of the magnetic polaron we solve the many-body Schrödinger equation of the underlying multicomponent system using the variational ML-MCTDHX method [42]. Specifically, a two-step truncation scheme is performed in the many-body wave function. First, the impurity-medium correlations are taken into account by expanding the many-body wave function |Ψ(t)⟩ with respect to D orthonormal species functions, i.e. with k = 1, 2, ..., D [42] for each component σ = B, I. Namely, we use the following truncated Schmidt decomposition
with the expansion coefficients λk known as the Schmidt weights and corresponding to the eigenvalues of the σ-component reduced density matrix. The latter is , with . A pre-requisite for the system to be entangled [76], or otherwise impurity-medium correlations are present, is that at least two different λk are populated.
Next, we express each of the above-described species functions as a linear superposition of time-dependent number-states with time-dependent expansion coefficients, ,
A particular number state corresponds to a permanent. It is constructed by dσ time-dependent variationally optimized single-particle functions (SPFs) , where l = 1, 2, ..., dσ and nl denote their occupation numbers. At this stage of the wave function truncation we account for intracomponent correlations.
Finally, the SPFs are expanded on a time-independent primitive basis. The latter is an dimensional discrete variable representation (DVR) for the impurity denoted by . For the spinor bosonic medium it is the tensor product (,) of the DVR basis for the spatial degrees-of-freedom and the two-dimensional pseudospin-1/2 basis {|↑⟩, |↓⟩}. Particularly, a SPF of the medium takes the spinor wave function form
with and being the time-dependent expansion coefficients of the spin-↑, ↓ respectively, see also references [26, 27] for further details.
To calculate the ground state of the underlying (N↑ + N↓ + 1)-body wave function describing the Hamiltonian of equation (1) we solve the respective ML-MCTDHX equations of motion [42] within the imaginary time propagation method. These equations are found upon applying, for instance, the Dirac–Frenkel variational principle [77] for the ansatz of equations (D1)–(D3). This process leads to a coupled set of D2 linear differential equations of motion for the λk (t) coefficients as well as and dB + dI nonlinear integrodifferential equations for the species functions and the SPFs respectively. The orbital configuration space C = (D; dB ; dI ) assigns the Hilbert space truncation. Here, the weakly interacting spinor Bose gas has a mesoscopic atom number resulting in suppressed intracomponent correlations which can be captured by a relatively small number of orbitals, here dB < 4. For the impurity we are able to employ a substantially larger orbital number, herein dI = 8, in order to describe strong impurity-medium correlations. As such, it is feasible to numerically solve the ML-MCTDHX equations of motion.
Footnotes
- 8
Notice that the term ∝δ in equation (2) breaks the Sx symmetry.
- 9
Deviations from the fully polarized (product) state occur due to the development of spin–spin host correlations. Nevertheless, the polarization lies along the x spin axis.
- 10
As we will argue below the dominant contribution of impurity-medium correlations stems from the coupling of the spin and spatial degrees-of-freedom in the many-body wave function, see section 4.4. Spatial medium correlations are suppressed in our setup for weak boson–boson interactions.
- 11
Notice that the magnetic Bose polaron can be also stable for gI↑ > gmag c despite possessing larger energy than the phase separated ground state. The coexistence of two stable phases within the same interaction interval indicates a hysteresis phenomenon with respect to ΩR referring to the presence or absence of the magnetic polaron.
- 12
Notice that in the case uα (xα − xI ) ≠ 0, and hold.
- 13
A fully ferromagnetic gas has for every N. The minimal value of does not allow for a perfectly anti-ferromagnetic configuration, since . Indeed, it is impossible that every pair of spins is anti-oriented which suppresses the perfect anti-ferromagnetic order.
- 14
μB , μI correspond to the solution of two algebraic equations, namely ∫dx|ψ↑(x; μB , μI )|2 + |ψ↓(x; μB , μI )|2 = N and ∫dx|ψI (x; μB , μI )|2 = 1, where ψσ (x; μB , μI ) is the wave function obtained from equation (16) as a function of μB , μI .
- 15
Namely, we keep constant the healing length of the medium, i.e. (equation (B7)), while simultaneously increasing the medium density, nB (0) ∝ μB /(g↑↓ + g). This is achieved by fixing gαα' N = const while increasing N and NI by the same factor.
- 16