This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Quantum-Fluctuation-Driven Dynamics of Droplet Splashing, Recoiling and Deposition in Ultracold Binary Bose Gases

Yinfeng Ma Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China    Xiaoling Cui [email protected] Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China Songshan Lake Materials Laboratory, Dongguan, Guangdong 523808, China
Abstract

Droplet impact on a surface is practically relevant to a variety of fields in nature and industry, while a complete control of its outcomes remains challenging due to various unmanageable factors. In this work, we propose the quantum simulation of droplet impact outcomes in the platform of ultracold atoms. Specifically, we study the quantum-fluctuation-driven dynamics (QFDD) of two-dimensional Bose-Bose mixtures from an initial Townes soliton towards the formation of a quantum droplet. By tuning the fluctuation energy of the initial Townes state through its size and number, the subsequent QFDD can produce various outcomes including splashing, recoiling, and deposition, similar to those in droplet impact dynamics. We have utilized the Weber number to identify the thresholds of splashing and recoiling, and further established a universal scaling law between the maximum spreading factor and the Weber number in the recoiling regime. In addition, we show that the residual QFDD in the deposition regime can be used to probe the collective breathing modes of a quantum droplet. Our results reveal a mechanism for the droplet impact outcomes, which can be directly tested in cold-atom experiments and can pave the way for exploring intriguing droplet dynamics in a clean and fully controlled quantum setting.

I Introduction

Given the wide practical relevance to both nature and industry [1, 2, 3, 4, 5, 6, 7], droplet impact dynamics on a surface has attracted much attention ever since the first study by Washington [8, 9]. Various impact outcomes-including splashing, receding/recoiling, rebound, and deposition-have been observed successfully in experiments [10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25]. In general, these dynamics were characterized by two physical observables, namely, the maximum spreading factor (β\beta) [12, 16, 23] and the splashing threshold(𝒦{\cal K}) [11, 13, 14, 15, 17, 18, 19, 20, 21, 22, 25], which were shown not only to depend on the properties of the droplet itself (size, density, surface tension, viscosity, impact velocity), but also to be strongly influenced by the surface condition(roughness, wettability) and surrounding gas (pressure, composition). Because of the complexities associated with various unmanageable factors, it is extremely challenging to deterministically parametrize β,𝒦\beta,\ {\cal K} and fully control the impact outcomes. In this situation, a common practice is to assume an ideal droplet impact (on a smooth solid surface at atmospheric condition) and then to quantify the actual dynamics by the Weber and Reynolds numbers, which, respectively, describe the relative strength of droplet inertia with respect to capillary and viscous forces [26, 27]. Various scaling laws between β,𝒦\beta,\ {\cal K} and these numbers have been proposed in the literature [26, 27], based on different models or empirical fitting from experimental data.

In the past few decades, ultracold atoms have emerged as an ideal platform for quantum simulation, given their extremely clean environment and the high controllability on the species, number, dimension, interaction strength etc [28, 29]. In particular, a recent important achievement in this field was the realization of a quantum droplet in both dipolar gas [30, 31, 32, 33, 34, 35, 36] and alkali bosonic mixtures [37, 38, 39, 40, 41, 42], with extremely dilute densities (10141015cm3\sim 10^{14}-10^{15}{\rm cm}^{-3}) that can be eight orders of magnitude lower than water. In forming these gaseous droplets, quantum fluctuations play an essential role in providing the repulsive force for their stabilization, for which they are called quantum droplets [43]. To date, the idea of a quantum droplet has been successfully extended to various atomic systems, including low dimensional ones [44, 45, 46, 47, 48, 49], Bose-Fermi mixtures [50, 51, 52, 53, 54, 55] and multi-component dipolar or alkali atomic mixtures [56, 57, 58]. The non-equilibrium properties of quantum droplets have also been investigated in terms of their dynamical formations [59, 60, 61] and collisions [62, 63]. These developments offer an unprecedented opportunity for simulating droplet impact dynamics in ultracold atoms, particularly, at the microscopic quantum level and in a highly controllable manner.

In this work, we demonstrate the capability of using ultracold Bose gases to simulate the droplet impact outcomes in a fully controlled quantum setting. Contrary to the conventional droplet impact setup, here there is no impact surface for the droplet, and the driving force of its dynamics is purely from its intrinsic energy due to quantum fluctuations. Specifically, we study the dynamical property of a two-dimensional (2D) Bose-Bose mixture with repulsive intraspecies and attractive inter-species couplings, whose ground state is a quantum droplet. To highlight the quantum effect in the dynamics, we have chosen the initial state as the Townes soliton generalized from the single-species case [64], which features a zero mean-field energy with continuous scale invariance as recently confirmed in experiments [65, 66, 67, 68]. In this way, the dynamics here is purely driven by quantum fluctuations and thus can be called quantum-fluctuation-driven dynamics (QFDD). It is found that by tuning the fluctuation energy of the initial Townes state through its size(σ0\sigma_{0}) and number(NN), the subsequent QFDD can produce various outcomes, including splashing, recoiling and deposition, similar to those in droplet impact dynamics. We have mapped out the dynamical phase diagram in the (σ0,N\sigma_{0},N) plane and employed the Weber number to characterize different phases. The splashing and recoiling thresholds are identified, and a universal scaling law is established between the maximum spreading factor and the Weber number for the recoiling dynamics, which is applicable for a considerably large parameter regime. Finally, we show that the long-time QFDD in the deposition regime can be used to extract the collective breathing modes of quantum droplet. These results can be directly tested in the current cold atoms experiments. In the Appendixes, we provide more details on the derivation of the generalized Townes soliton with unequal masses, as well as on the numerical simulations.

II Model

We start from the energy functional of two-species bosons in 2D (=1\hbar=1 for brevity):

(𝝆)\displaystyle\cal{E}(\bm{\rho}) =\displaystyle= i=1,2ϕi(𝝆)𝝆22mϕi(𝝆)+ijgij2ni(𝝆)nj(𝝆)\displaystyle-\sum_{i=1,2}\phi_{i}^{*}(\bm{\rho})\frac{\nabla_{\bm{\rho}}^{2}}{2m}\phi_{i}(\bm{\rho})+\sum_{ij}\frac{g_{ij}}{2}n_{i}(\bm{\rho})n_{j}(\bm{\rho}) (1)
+LHY(ni(𝝆));\displaystyle+{\cal E}_{\rm LHY}(n_{i}(\bm{\rho}));

here 𝝆=(x,y)\bm{\bm{\rho}}=(x,y) is the 2D coordinate; ϕi\phi_{i} is the wavefunction of the ii-th species and ni=|ϕi|2n_{i}=|\phi_{i}|^{2} is its density; gijg_{ij} is the bare coupling between ii- and jj-species, which can be expressed as gij=4πaij/(mlz)g_{ij}=4\pi a_{ij}/(ml_{z}) in quasi-2D geometry, with aija_{ij} the s-wave scattering length and lzl_{z} the characteristic length along the confined (zz) direction; and LHY{\cal E}_{\rm LHY} is the Lee-Huang-Yang(LHY) correction from quantum fluctuations, and for quasi-2D bosons near the mean-field instability point (δaa12+a11a220\delta a\equiv a_{12}+\sqrt{a_{11}a_{22}}\sim 0) it reads [44, 47]

LHY=η2ln(ηe)8πmlz4,withη=4πlz(a11n1+a22n2).{\cal E}_{\rm LHY}=\frac{\eta^{2}\ln(\eta\sqrt{e})}{8\pi ml_{z}^{4}},\ \ \ \ {\rm with}\ \ \eta=4\pi l_{z}(a_{11}n_{1}+a_{22}n_{2}). (2)

It has been shown that this LHY term can balance with the mean-field force and result in a self-bound droplet as the ground state [44]. Given the n2ln(n)n^{2}\ln(n) dependence of LHY energy, the 2D droplet can exist at both mean-field collapse and stable regimes [44] and with any infinitesimal atom number [49]. At a sufficiently large number, the 2D droplet develops a flat-top structure in its density profile, similar to the 3D case [43].

Given Eq.(1), the dynamics of {ϕi}\{\phi_{i}\} is governed by the time-dependent Gross-Pitaevskii(GP) equations:

itϕi=(𝝆22m+jgijnj+LHYni)ϕi.i\partial_{t}\phi_{i}=\left(-\frac{\nabla_{\bm{\rho}}^{2}}{2m}+\sum_{j}g_{ij}n_{j}+\frac{\partial{\cal E}_{LHY}}{\partial n_{i}}\right)\phi_{i}. (3)

In this work, we will focus on the solution with zero angular momentum, since the ground state and the Townes soliton both stay within this sector, and different angular momentum sectors are decoupled from each other. In this sector, we can replace the coordinate 𝝆\bm{\bm{\rho}} simply by its magnitude ρ=|𝝆|\rho=|\bm{\rho}|. More details of solving Eq.(3) in the discretized coordinate and time space have been given in the Appendixes.

Throughout the paper, we specifically consider the two hyperfine states of 39K atoms, |1|F=1,mF=1,|2|F=1,mF=0|1\rangle\equiv|F=1,m_{F}=-1\rangle,\ |2\rangle\equiv|F=1,m_{F}=0\rangle, as has been well studied in the droplet experiments [37, 38, 39]. In this case, a11=33.5aB,a12=53aBa_{11}=33.5a_{B},\ a_{12}=-53a_{B} (aBa_{B} is the Bohr radius), and a22a_{22} is highly tunable by magnetic field. The confinement length is chosen as lz=0.08μml_{z}=0.08\mu m.

III Generalized Townes soliton

For spinless bosons in 2D, it is known that the kinetic term and mean-field attraction can support a special stationary solution called the Townes soliton [64]. This state features zero energy and continuous scale invariance, and it can only exist when the boson number and coupling strength satisfy N|g|=5.85/mN|g|=5.85/m. Such a special solution has been successfully observed in both non-linear optics [65] and ultracold atoms [66, 67, 68]. In these experiments, the LHY correction takes little effect as it is much smaller than the mean-field part. In the following, we will show that the Townes soliton can be generalized to two-species bosons if equally neglecting the LHY correction.

By omitting the LHY term in (3), we can see that the two GP equations for {ϕ1,ϕ2}\{\phi_{1},\phi_{2}\} can support a single-mode solution ϕi=Niϕexp(iμit)\phi_{i}=\sqrt{N_{i}}\phi\exp(-i\mu_{i}t) as long as

N1N2=g22g12g11g12,\frac{N_{1}}{N_{2}}=\frac{g_{22}-g_{12}}{g_{11}-g_{12}}, (4)

where μ1=μ2μ\mu_{1}=\mu_{2}\equiv\mu, and the single mode ϕ\phi satisfies:

(𝝆22m+Ngeff|ϕ|2)ϕ=μϕ.\left(-\frac{\nabla_{\bm{\rho}}^{2}}{2m}+Ng_{\rm eff}|\phi|^{2}\right)\phi=\mu\phi. (5)

Here N=N1+N2N=N_{1}+N_{2} is the total number, and the effective interaction geffg_{\rm eff} is given by

geff=g11g22g122g11+g222g12.g_{\rm eff}=\frac{g_{11}g_{22}-g_{12}^{2}}{g_{11}+g_{22}-2g_{12}}. (6)

Apparently we have geff<0g_{\rm eff}<0 in the mean-field collapse regime (g12<g11g22g_{12}<-\sqrt{g_{11}g_{22}}). It is then straightforward to check that under the condition

Nm|geff|=5.85,Nm|g_{\rm eff}|=5.85, (7)

there exists a sequence of zero-energy eigenstates with continuous scale invariance, i.e., the eigenstate nature and zero-energy property will not change under an arbitrary scaling transformation ϕ(𝝆)λϕ(λ𝝆)\phi(\bm{\rho})\rightarrow\lambda\phi(\lambda\bm{\rho}) (accordingly Eλ2EE\rightarrow\lambda^{2}E). These stationary solutions are the generalized Townes soliton for two-species bosons. Note that a similar generalization also works for the case of unequal masses, with slight modifications in Eqs.(4,6) as shown in the Appendixes.

Refer to caption

Figure 1: (Color online) Test of generalized Townes soliton for two-species bosons (neglecting the LHY correction). (a) A stationary Townes profile when the total number NN and number ratio N1/N2N_{1}/N_{2} satisfy (7) and (4) simultaneously. (b) Density profile at different times when the total number is changed to 0.8N0.8N, while the number ratio is the same as in (a). (c) Density profile at different times when the total number is 1.2N1.2N, with the same number ratio as in (a). (d) Density profile at different times with the same total number NN as in (a) while keeping N1/N2=1N_{1}/N_{2}=1. Here we take two-species 39K atoms with a11=33.5aB,a12=53aBa_{11}=33.5a_{B},\ a_{12}=-53a_{B} and a22=83.1aBa_{22}=83.1a_{B}, which gives N=6500N=6500 for a stationary Townes state according to Eqs.(6,7); the initial size of the Townes profile is taken as 7μm7\mu m.

In figure 1(a), we confirm the stationary Townes soliton for two-species bosons once the total number NN satisfies (7) and the number ratio N1/N2N_{1}/N_{2} satisfies (4). In comparison, if we change NN to be smaller or larger, the original profile will shrink (figure 1(b)) or expand (figure 1(c)) with time. The profile is also unstable if N1/N2N_{1}/N_{2} deviates from (4), see figure 1(d). In a word, both conditions (7) and (4) are required in supporting a stationary two-species Townes solution.

IV Quantum-fluctuation driven dynamics

A crucial difference between the single- and two-species bosons is that quantum fluctuations play an important role in the latter, which can lead to droplet formation as a ground state. It then follows that starting from the generalized Townes soliton, which is the mean-field stationary solution for two-species bosons, the quantum fluctuation can destabilize it strongly and drive its time evolution towards the droplet formation. Such dynamics can be called the quantum-fluctuation driven dynamics (QFDD), also in light of the fact that the total energy of Townes soliton is purely given by the LHY part, E(t=0)=ELHYE(t=0)=E_{\rm LHY}.

Refer to caption

Figure 2: (Color online) Quantum-fluctuation driven dynamics(QFDD) of deposition (b1), recoiling (b2) and splashing (b3) starting from initial Townes solitons with different sizes. The parameters of {aij}\{a_{ij}\} and initial NN are the same as in figure 1. (a) Total energy per particle (E/N=ELHY/NE/N=E_{\rm LHY}/N) for the initial Townes state as functions of its size (σ\sigma). The color on the curve denotes the corresponding Weber number. The dashed line shows the energy from Gaussian ansatz, and the lower dotted line shows the energy of ground state droplet. The energy unit is 1/(ml02)1/(ml_{0}^{2}) with l0=1μml_{0}=1\mu m. (b1-b3) Three typical QFDD starting from different initial Townes states, which respectively simulate the droplet impact outcomes of deposition (b1), recoiling (b2) and splashing(b3). The initial sizes of Townes solitons are respectively σ0(μm)=4.75(b1), 2.5(b2), 2(b3)\sigma_{0}(\mu m)=4.75(b1),\ 2.5(b2),\ 2(b3), as located by circles in (a). (c) Time evolution of the cloud size, σ(t)\sigma(t), for the dynamics in (b1-b3).

In figure 2(a), we show that the total energy (=ELHY=E_{\rm LHY}) of two-species Townes soliton can be conveniently tuned by its size σ\sigma, taking a typical combination of {N,gij}\{N,g_{ij}\} that satisfies (4,7). In particular, ELHYE_{\rm LHY} varies non-monotonically with σ\sigma and shows a minimum at certain finite σ\sigma. To understand this behavior, we employ a Gaussian ansatz to approximate the single mode ϕ(ρ)=exp[ρ2/(2σ2)]/(πσ)\phi(\rho)=\exp[-\rho^{2}/(2\sigma^{2})]/(\sqrt{\pi}\sigma), which leads to the total energy E=2πρ𝑑ρ(ρ)E=\int 2\pi\rho d\rho{\cal E}(\rho) as

E=N2mσ2+N2geff4πσ2+N2a¯2mlz2σ2ln4Na¯lzσ2,\begin{split}E=\frac{N}{2m\sigma^{2}}+\frac{N^{2}g_{\rm eff}}{4\pi\sigma^{2}}+\frac{N^{2}\bar{a}^{2}}{ml_{z}^{2}\sigma^{2}}\ln\frac{4N\bar{a}l_{z}}{\sigma^{2}},\end{split} (8)

with a¯a11a22\bar{a}\equiv\sqrt{a_{11}a_{22}}. We can see that the first two terms can support zero-energy states with arbitrary σ\sigma under the condition Nm|geff|=2πNm|g_{\rm eff}|=2\pi, which are just the simplified Gaussian version of the Townes soliton. However, when including the third LHY term, the total energy(=ELHY=E_{\rm LHY}) will deviate from zero, and the Townes profile is no longer stationary. Given the expression ELHYln(σ2)/σ2E_{\rm LHY}\sim-\ln(\sigma^{2})/\sigma^{2}, we can easily arrive at a non-monotonic ELHYσE_{\rm LHY}\sim\sigma dependence with energy minimum Emin=Na¯/(4lz3e)E_{\rm min}=-N\bar{a}/(4l_{z}^{3}e) at σmin=4Na¯lze\sigma_{\rm min}=\sqrt{4N\bar{a}l_{z}e}. As shown in figure 2(a), ELHYE_{\rm LHY} from a Gaussian ansatz provides a qualitatively good prediction to the EσE\sim\sigma lineshape of real Townes solutions.

Given the easily tunable fluctuation energy of the initial Townes state, the subsequent QFDD can exhibit rich dynamical outcomes. In figure 2(b1,b2,b3), we show the time evolution of density profiles for three typical QFDDs:

(I) Deposition. When the initial size σ0\sigma_{0} is close to σmin\sigma_{\rm min} and ELHYE_{\rm LHY} is small, the QFDD shows a typical deposition behavior (figure 2(b1)). Specifically, as time passes the system repels a small proportion of atoms outside, and the rest automatically follows the profile of a ground-state droplet with additional periodic oscillations. As discussed later, such residual oscillation can be used to probe the collective breathing modes of a quantum droplet.

(II) Recoiling. As σ0\sigma_{0} deviates more from σmin\sigma_{\rm min} and ELHYE_{\rm LHY} gets larger, the system enters the recoiling regime. As shown in figure 2(b2), at early times a considerable portion of atoms are repelled outside, while at some point they stop to spread and flow back to merge with the central part. Such back-flow (or recoiling) can be attributed to the competition between the surface tension and the kinetic energy of the cloud during the dynamics.

(III) Splashing. When σ0\sigma_{0} deviates significantly from σmin\sigma_{\rm min} and ELHYE_{\rm LHY} is large enough, the system shows a rapid splashing dynamics, see figure 2(b3). In this case, the large ELHYE_{\rm LHY} causes a drastic change of the initial Townes profile in a short time, i.e., the cloud quickly splits into two pieces. In this process, ELHYE_{\rm LHY} converts to the large kinetic energy of the outgoing part, such that it completely separates from the central part and flows away forever.

The above dynamics can be well distinguished by monitoring the mean size of the dynamical system, σ(t)=𝝆2t\sigma(t)=\sqrt{\langle\bm{\rho}^{2}\rangle_{t}}. As shown in figure 2(c), at longer times σ(t)\sigma(t) is almost static in the deposition regime, while it shows a non-monotonic behavior in the recoiling regime and a continuous increase in the splashing regime.

To this end, we have shown that the QFDD starting from the Townes states can produce rich dynamical phases, including the deposition, recoiling, and splashing, which perfectly mimic the droplet impact outcomes as studied in the literature [10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25]. However, different from these existing studies of droplet impact on a surface, in our case the driving force of the dynamics is purely from the intrinsic energy contributed by quantum fluctuations. This introduces a mechanism for these fluid dynamics. Meanwhile, in QFDD there are no complexities caused by the impact surface or environment, and the dynamical outcome can be fully controlled by adjusting the size (σ0)(\sigma_{0}) and the number (N)(N) of the initial state. Therefore, the ultracold atoms provide an extremely clean and convenient platform to simulate droplet dynamics, where the quantum effect can be well manipulated and the fluid mechanics can be understood in a more deterministic way.

V Weber number and dynamical phase diagram

We now quantify various dynamical phases in QFDD by the Weber number. Note that the Reynolds number is irrelevant here because of the zero viscosity of Bose condensates. In conventional droplet impact dynamics [26, 27], the Weber number is defined as W=ρdDv2/γW=\rho_{d}Dv^{2}/\gamma, where ρd,D,v,γ\rho_{d},\ D,\ v,\ \gamma respectively denote the droplet density, diameter, impact velocity, and surface tension. It measures the relative strength of droplet inertia with respect to capillary force. Here, we generalize the definition of WW to describe the quantum dynamics in general:

W=ΔEDγ,W=\frac{\Delta E}{D\gamma}, (9)

where ΔE\Delta E is the energy difference between the initial state (here the Townes soliton) and the true ground state given the same initial parameters (N,gijN,\ g_{ij}); DD and γ\gamma are the same as before, i.e., the droplet diameter and surface tension. Specifically, we have γ=𝑑𝝆((𝝆)μ1|ϕ1|2μ2|ϕ2|2)\gamma=\int d\bm{\bm{\rho}}\left({\cal E}(\bm{\bm{\rho}})-\mu_{1}|\phi_{1}|^{2}-\mu_{2}|\phi_{2}|^{2}\right), with μi\mu_{i} the chemical potential for the i-th species.

As shown in the color plot in figure 2(a), WW defined in (9) can well characterize different dynamical phases in QFDD. Namely, a low WW corresponds to the deposition dynamics, where the small ΔE\Delta E can be well absorbed by the surface change of the droplet; as WW increases, the system enters the recoiling regime and finally end up at splashing, where the large ΔE\Delta E overwhelms the capacity of the droplet surface and causes it to change drastically. Hereafter, we refer to the critical WW at the recoiling-splashing boundary as the splashing threshold (𝒦s{\cal K}_{s}), and that at the deposition-recoiling boundary as the recoiling threshold (𝒦r{\cal K}_{r}).

Refer to caption

Figure 3: (Color online) Characterization of different dynamical phases and their boundaries in-between. (a) Contour plot of Weber number in the parameter plane of initial Townes size (σ0\sigma_{0}) and number(NN). The splashing, recoiling and deposition phases are respectively marked by ’S’, ’R1,2’ and ’D’ in the diagram. The white squares with white dashed lines mark their boundaries. Gray dashed line shows the location of energy minimum for initial Townes state. (b) Splashing threshold (𝒦s{\cal K}_{s}) and two recoiling thresholds (𝒦r1,𝒦r2{\cal K}_{r1},{\cal K}_{r2}) as functions of NN along the phase boundaries in (a). Note that when changing NN, the coupling constant a22a_{22} of 39K atoms also changes according to the constraint (6,7) for initial Townes states. The other couplings {a11,a12}\{a_{11},a_{12}\} are the same as in figure 1.

In figure 3(a), we map out the dynamical phase diagram in the (σ0,N\sigma_{0},N) parameter plane. Distinct dynamical outcomes of QFDD, including deposition (’D’), recoiling (’R’) and splashing (’S’), are identified by monitoring the mean size σ(t)\sigma(t) of the cloud during expansion (see figure 2(c)). In addition, we show the contour plot of WW in the (σ0,N\sigma_{0},N) plane, and one can see clearly that the ’D’, ’R’ and ’S’ phases respectively correspond to the small, intermediate and large WW regions. Due to the non-monotonic dependence of WW on σ0\sigma_{0} (as shown in figure 2(a)), there are two recoiling regions in the diagram, as marked by ’R1’ and ’R2’. In figure 3(b), we extract the two recoiling thresholds(𝒦r1,𝒦r2{\cal K}_{r1},{\cal K}_{r2}) and the splashing threshold(𝒦s{\cal K}_{s}) along the phase boundaries as varying NN. One can see that 𝒦r1𝒦r2{\cal K}_{r1}\approx{\cal K}_{r2} are given by a constant 0.12\sim 0.12, while 𝒦s{\cal K}_{s} is a much larger value and continuously increases with NN.

VI Maximum spreading factor

Another important physical quantity to characterize the droplet impact dynamics is the maximum spreading factor β=σmax/σ0\beta=\sigma_{\rm max}/\sigma_{0}, as defined by the ratio between the maximum spreading radius (σmax\sigma_{\rm max}) and the initial one (σ0\sigma_{0}). Clearly one has β1\beta\sim 1 for the deposition dynamics and β\beta\rightarrow\infty for splashing. An interesting behavior of β\beta shows up in the recoiling regime, where β\beta is finite and varies sensitively with WW.

Refer to caption

Figure 4: (Color online) Universal scaling between the maximum spreading factor (β\beta) and the Weber number (WW) in the recoiling regime. Here we take four given atom numbers of the initial Townes state with tunable sizes, as following the trajectory along the horizontal lines in Fig.3(a). The discrete points are from numerical calculations and the lines are fitting functions according to Eq.(10).

In figure 4, we extract β\beta as a function of WW along the horizontal lines in figure 3(a), i.e., by varying σ0\sigma_{0} at several fixed NN in the recoiling regime. Apart from the region near the phase boundaries (β1\beta\rightarrow 1 or large enough), the data of β\beta for any given NN well follow the scaling relation:

β1Wα,withα=1.59.\beta-1\propto W^{\alpha},\ \ \ {\rm with}\ \ \alpha=1.59. (10)

As shown in figure 4, such scaling works well for a wide parameter regime with N/104(0.5,2)N/10^{4}\in(0.5,2), β(1.5,10)\beta\in(1.5,10) and W(0.2,1)W\in(0.2,1). Therefore, we expect the scaling law in (10) to reflect a very robust intrinsic property of the ultracold fluid during the recoiling QFDD.

VII Breathing modes

Finally, we demonstrate that the deposition regime of QFDD can be used to probe the breathing modes of quantum droplet. The breathing mode can be theoretically obtained as follows. Assuming a small fluctuation mode δϕi\delta\phi_{i} for the i-th species boson, and only keeping the lowest-order fluctuations in the GP equations(3), we obtain the following equations for {δϕi}\{\delta\phi_{i}\}:

itδϕ1=(𝝆22m+g11n1+g12n2+LHYn1)δϕ1+g11n1(δϕ1+δϕ1)+g12ϕ1ϕ2(δϕ2+δϕ2)+2LHYn12n1(δϕ1+δϕ1)+2LHYn1n2ϕ1ϕ2(δϕ2+δϕ2),\begin{split}i\partial_{t}\delta\phi_{1}&=\left(-\frac{\nabla^{2}_{\bm{\bm{\rho}}}}{2m}+g_{11}n_{1}+g_{12}n_{2}+\frac{\partial{\cal E}_{LHY}}{\partial n_{1}}\right)\delta\phi_{1}+g_{11}n_{1}(\delta\phi_{1}+\delta\phi_{1}^{*})+g_{12}\phi_{1}\phi_{2}(\delta\phi_{2}+\delta\phi_{2}^{*})\\ &+\frac{\partial^{2}{\cal E}_{LHY}}{\partial n_{1}^{2}}n_{1}(\delta\phi_{1}+\delta\phi_{1}^{*})+\frac{\partial^{2}{\cal E}_{LHY}}{\partial n_{1}\partial n_{2}}\phi_{1}\phi_{2}(\delta\phi_{2}+\delta\phi_{2}^{*}),\\ \end{split} (11)
itδϕ2=(𝝆22m+g12n1+g22n2+LHYn2)δϕ2+g22n2(δϕ2+δϕ2)+g12ϕ1ϕ2(δϕ1+δϕ1)+2LHYn1n2ϕ1ϕ2(δϕ1+δϕ1)+2LHYn22n2(δϕ2+δϕ2).\begin{split}i\partial_{t}\delta\phi_{2}&=\left(-\frac{\nabla^{2}_{\bm{\bm{\rho}}}}{2m}+g_{12}n_{1}+g_{22}n_{2}+\frac{\partial{\cal E}_{LHY}}{\partial n_{2}}\right)\delta\phi_{2}+g_{22}n_{2}(\delta\phi_{2}+\delta\phi_{2}^{*})+g_{12}\phi_{1}\phi_{2}(\delta\phi_{1}+\delta\phi_{1}^{*})\\ &+\frac{\partial^{2}{\cal E}_{LHY}}{\partial n_{1}\partial n_{2}}\phi_{1}\phi_{2}(\delta\phi_{1}+\delta\phi_{1}^{*})+\frac{\partial^{2}{\cal E}_{LHY}}{\partial n_{2}^{2}}n_{2}(\delta\phi_{2}+\delta\phi_{2}^{*}).\\ \end{split} (12)

According to the standard Bogoliubov analysis, we search for solutions of the form:

δϕi=exp(iμit)j(uij(𝝆)exp(iωjt)+vij(𝝆)exp(iωjt)).\delta\phi_{i}=\exp(-i\mu_{i}t)\sum_{j}\left(u_{ij}(\bm{\rho})\exp(-i\omega_{j}t)+v_{ij}^{*}(\bm{\rho})\exp(i\omega_{j}t)\right). (13)

Here ωj\omega_{j} is the jj-th collective (eigen-)mode of the system. These modes can be extracted from the following coupled equations for uij(𝝆)u_{ij}(\bm{\rho}) and vij(𝝆)v_{ij}(\bm{\rho}):

(L1+M1M12M1M12M12L2+M2M12M2M1M12(L1+M1)M12M12M2M12(L2+M2))(u1ju2jv1jv2j)=ωj(u1ju2jv1jv2j),\left(\begin{matrix}L_{1}+M_{1}&M_{12}&M_{1}&M_{12}\\ M_{12}&L_{2}+M_{2}&M_{12}&M_{2}\\ -M_{1}&-M_{12}&-(L_{1}+M_{1})&-M_{12}\\ -M_{12}&-M_{2}&-M_{12}&-(L_{2}+M_{2})\end{matrix}\right)\left(\begin{matrix}u_{1j}\\ u_{2j}\\ v_{1j}\\ v_{2j}\end{matrix}\right)=\omega_{j}\left(\begin{matrix}u_{1j}\\ u_{2j}\\ v_{1j}\\ v_{2j}\end{matrix}\right), (14)

where

Li=𝝆22m+jgij|ϕj|2+LHYniμiMi=gii|ϕi|2+2LHYni2niM12=g12ϕ1ϕ2+2LHYn1n2ϕ1ϕ2.\begin{split}L_{i}&=-\frac{\nabla^{2}_{\bm{\bm{\rho}}}}{2m}+\sum_{j}g_{ij}|\phi_{j}|^{2}+\frac{\partial{\cal E}_{LHY}}{\partial n_{i}}-\mu_{i}\\ M_{i}&=g_{ii}|\phi_{i}|^{2}+\frac{\partial^{2}{\cal E}_{LHY}}{\partial n_{i}^{2}}n_{i}\\ M_{12}&=g_{12}\phi_{1}\phi_{2}+\frac{\partial^{2}{\cal E}_{LHY}}{\partial n_{1}\partial n_{2}}\phi_{1}\phi_{2}.\\ \end{split} (15)

In Fig.5, we show the lowest two breathing modes (ω1,ω2\omega_{1},\omega_{2}) below the atom emission threshold (μ1,μ2-\mu_{1},-\mu_{2}), see solid lines. In our numerics, we have obtained ωj\omega_{j} by discretizing the coordinate space and exactly diagonalizing the resulted large matrix in the left-hand-side of (14). Note that the breathing modes, by definition, stay in the same zero angular momentum sector as the Townes state and the ground state droplet. More details on the numerics are given in the Appendixes.

Refer to caption

Figure 5: (Color online) Extracting collective breathing modes from QFDD in the deposition regime. Here we take a22=83.8aBa_{22}=83.8a_{B} and the other parameters are the same as in Figure 1; the initial atom number is N(104)=7.3N(10^{4})=7.3. (a) and (b) show the boson densities at different locations 𝝆=0\bm{\rho}=0 (center), 16.4μm16.4\mu m in the time and frequency domains (related by Fourier transformation), staring from an initial Townes state with size σ0=18μm\sigma_{0}=18\mu m. The two red arrows in (a) mark the time interval for Fourier transformation. (c) Collective excitation spectrum as a function of droplet number NdN_{d}. The lines are theoretical results based on Bogoliubov analysis, and the discrete data with error bar are extracted from the real density oscillations in the deposition dynamics. In the latter case the residual number NdN_{d} can be effectively tuned by setting different sizes of initial Townes state, and here we take two sizes σ0(μm)=16, 18\sigma_{0}(\mu m)=16,\ 18 which lead to Nd(104)=6.5, 6.2N_{d}(10^{4})=6.5,\ 6.2 respectively. The energy unit is 1/(ml02)1/(ml_{0}^{2}) with l0=1μml_{0}=1\mu m.

In figure 5, we show how to extract collective breathing modes from the QFDD in the deposition regime for a given set of coupling strengths {aij}\{a_{ij}\} and initial atom number NN. Figure 5(a) shows the typical density oscillations with time at different locations. One can see that such oscillations are synchronic for both species and at different locations, and therefore their Fourier transformations give the same peak frequencies; see figure 5(b). In figure 5(c), we plot out two sets of extracted peak frequencies (ω1,ω2\omega_{1},\omega_{2}) with different residual droplet numbers NdN_{d}. Here NdN_{d} can be tuned effectively by setting different sizes of the initial Townes state. For all the extracted data we have collected 30 samples by varying the real-space locations or the time intervals in Fourier transformation, from which we obtain both the mean value and the variance. One can see that the extracted results match very well with theoretical predictions of collective breathing modes from Eq.14 (solid lines).

We note that a previous study has extracted one branch of a collective mode from the formation dynamics of a quantum droplet [59], while the dynamics there is not QFDD. Here, we have shown in figure 5 that two branches of collective modes can be simultaneously extracted from the residual oscillations in QFDD. Moreover, we have checked that by taking different {aij}\{a_{ij}\} and initial NN, one can also get access to the other regime with only a single breathing mode. In comparison to the single-mode case, here the main complexity brought by the presence of two modes is the appearance of additional frequency peaks at ω2±ω1\omega_{2}\pm\omega_{1} due to their interference. To correctly identify the breathing modes {ωj}\{\omega_{j}\} from the multiple frequency peaks, it is important to note that these modes are associated with the most pronounced peaks among all others. This can be understood by analyzing the spin densities ni=|ϕi+δϕi|2n_{i}=|\phi_{i}+\delta\phi_{i}|^{2}, with δϕi\delta\phi_{i} given in Eq.(13). Obviously, one can see that the Fourier transform of nin_{i} results in multiple peaks at {ωj}\{\omega_{j}\}, {2ωj}\{2\omega_{j}\} and {ωi±ωj}\{\omega_{i}\pm\omega_{j}\}, while the ones at {ωj}\{\omega_{j}\} have the largest magnitudes, which depend linearly on the strength of the fluctuation mode (the others all have a quadratic dependence). Based on this principle, one can easily identify the correct breathing modes (ω1,ω2\omega_{1},\omega_{2}) as marked in figure 5(b).

VIII Experimental relevance

Our results can be readily tested in the current cold-atom experiments. The initial Townes profile with a given amplitude and size can be imprinted by properly designing the optical potential applied to the atoms, as successfully implemented in previous experiments on two-component Bose gases [68, 69]. The subsequent dynamics of the system can then be measured through the in situ image, and various dynamical outcomes can be distinguished typically within tends of milliseconds, as shown in Fig.2. Within this time scale severe atom loss in the 39K droplet can be effectively avoided [37]. On the other hand, for long time dynamics, the atom loss will play an essential role and may even impede the breathing mode detection, which requires a typical time scale of hundreds of milliseconds, as shown in Fig.5(a1,a2). For that, one can resort to the heteronuclear mixtures of 41K and 87Rb, whose lifetime has been shown to be much longer (1s\sim 1s) due to the low droplet density therein to suppress the atom loss [40, 41]. Since the Townes soliton can be well extended to boson mixtures with unequal masses (see the Appendixes), and the controllability of LHY energy by the size and number generally applies to these systems, we expect that various dynamical outcomes and the breathing mode extraction can be tested equally in heteronuclear Bose gases.

IX Summary and discussion

In summary, we have demonstrated the quantum simulation of droplet impact outcomes in ultracold boson mixtures. Various dynamical phases including splashing, recoiling, and deposition, have been revealed. A remarkable difference here is that these dynamics are purely driven by quantum fluctuations, instead of the mechanical impact force in previous studies. Given the easy manipulation of initial states with tunable fluctuation energies, the current cold-atom platform provides complete control of these dynamics in the microscopic quantum level. To characterize different dynamics, we have introduced the Weber number and examined two important physical quantities, namely, the splashing/recoiling threshold and the maximum spreading factor. We have also proposed to extract the collective breathing modes from the residual dynamics in the deposition regime. These results are directly relevant to ongoing cold-atom experiments.

Furthermore, we remark that our work is in distinct contrast to the previous studies of QFDD in ultracold atoms, where a visible dynamics can only be achieved for small condensates [70, 71, 72, 73]. This is because usually the energy difference per particle between the mean-field and the true quantum ground states decays rapidly as the number NN increases [74, 75, 76, 77] , and thus for large NN the quantum fluctuation energy is negligibly small as compared to the mean-field one. However, this is not the case for a quantum droplet. For a static droplet, quantum fluctuations have been shown to provide an indispensable repulsive force for its stabilization, without which the whole system will collapse [43]. In parallel, our current work reveals the equally significant role of quantum fluctuations played in the non-equilibrium dynamics of a quantum dropet, regardless of whether the droplet is small or large.

Finally, we anticipate that the dynamical outcomes of splashing, recoiling, and deposition revealed in this work may be equivalent with other initial conditions when the dynamics is not purely driven by quantum fluctuations. In fact, the generalized Weber number as defined in Eq.(9) can be applied to any initial state, and the key issue is whether the additional energy of such a state (as compared to the true ground state) can be absorbed by the surface tension of a droplet during dynamics. In this sense, Eq.(9) may serve as a unified quantity to understand the different dynamics of quantum droplets with various initial conditions.

Acknowledgment

The work is supported by the National Key Research and Development Program of China (2018YFA0307600), the National Natural Science Foundation of China (12074419, 12134015, 12205365), and the Strategic Priority Research Program of Chinese Academy of Sciences (No. XDB33000000).

Appendix A Townes soliton for two-species bosons with unequal masses

Neglecting the LHY correction, we write down the Gross-Pitaevskii(GP) equation for two-species bosons with unequal mass (m1,m2m_{1},m_{2}):

itϕ1=(𝝆22m1+g11|ϕ1|2+g12|ϕ2|2)ϕ1;\displaystyle i\partial_{t}\phi_{1}=\left(-\frac{\nabla_{\bm{\rho}}^{2}}{2m_{1}}+g_{11}|\phi_{1}|^{2}+g_{12}|\phi_{2}|^{2}\right)\phi_{1}; (16)
itϕ2=(𝝆22m2+g12|ϕ1|2+g22|ϕ2|2)ϕ2.\displaystyle i\partial_{t}\phi_{2}=\left(-\frac{\nabla_{\bm{\rho}}^{2}}{2m_{2}}+g_{12}|\phi_{1}|^{2}+g_{22}|\phi_{2}|^{2}\right)\phi_{2}. (17)

In the case m1m2m_{1}\neq m_{2}, we define the mass-imbalance parameter as w=m2/m1w=m_{2}/m_{1}. By comparing the two GP equations, we can see that they can support a single-mode solution ϕi=Niϕexp(iμit)\phi_{i}=\sqrt{N_{i}}\phi\exp(-i\mu_{i}t) as long as

N1N2=wg22g12g11wg12,\frac{N_{1}}{N_{2}}=\frac{wg_{22}-g_{12}}{g_{11}-wg_{12}}, (18)

where μ1=wμ2\mu_{1}=w\mu_{2}, and the single mode ϕ\phi satisfies:

(𝝆22m1+Ngeff|ϕ|2)ϕ=μ1ϕ.\left(-\frac{\nabla_{\bm{\rho}}^{2}}{2m_{1}}+Ng_{\rm eff}|\phi|^{2}\right)\phi=\mu_{1}\phi. (19)

Here N=N1+N2N=N_{1}+N_{2} is the total number, and the effective interaction geffg_{\rm eff} is given by

geff=w(g11g22g122)g11+wg22(w+1)g12.g_{\rm eff}=\frac{w(g_{11}g_{22}-g_{12}^{2})}{g_{11}+wg_{22}-(w+1)g_{12}}. (20)

Again we have geff<0g_{\rm eff}<0 in the mean-field collapse regime (g12<g11g22g_{12}<-\sqrt{g_{11}g_{22}}), and the Townes solution occurs when

Nm1|geff|=5.85.Nm_{1}|g_{\rm eff}|=5.85. (21)

For the equal mass case (w=1w=1), these equations automatically reduce to Eqs.(4-7) in the main text.

Appendix B Numerical Method

In our numerical simulations, we have considered the solution with zero angular momentum. This choice is because the angular momentum is preserved by the Hamiltonian, and both the ground state and the Townes soliton stay exactly in the zero angular momentum sector. Moreover, the breathing mode, by definition, also stays in the zero angular momentum sector. Given this constraint, in the following we simply replace the coordinate 𝝆\bm{\bm{\rho}} by its magnitude 𝝆=|𝝆|\bm{\rho}=|\bm{\bm{\rho}}|.

In solving the GP equation of the form itϕ(ρ,t)=Hϕ(ρ,t)i\partial_{t}\phi(\rho,t)=H\phi(\rho,t), we have discretized the coordinate and time as ρj=jδ,tn=nτ\rho_{j}=j\delta,\ t_{n}=n\tau, where j,nj,n are all integers and we have taken small intervals δ=0.005μm,τ=2.74×103ms\delta=0.005\mu m,\ \tau=2.74\times 10^{-3}ms. Then starting from a given state ϕn\phi_{n} at time tnt_{n}, we obtained the new wavefunction ϕn+1\phi_{n+1} at the next time tn+1t_{n+1} as follows.

First, we split HH into the local and non-local parts: H=H1+H2H=H_{1}+H_{2}, where the derivatives of ρ\rho are all contained in H2H_{2}. In this case, we simply have H2H_{2} as the kinetic energy H2=12m[2ρ2+1ρρ]H_{2}=-\frac{1}{2m}\left[\frac{\partial^{2}}{\partial\rho^{2}}+\frac{1}{\rho}\frac{\partial}{\partial\rho}\right], and H1H_{1} is the rest part that solely depends on local densities.

As the first step of the update, the local H1H_{1} produces an intermediate state ϕ\phi^{\prime} from ϕn\phi_{n}:

ϕ=exp(iH1τ)ϕn.\phi^{\prime}=\exp(-iH_{1}\tau)\phi_{n}. (22)

Then, we perform the time evolution generated by H2H_{2} with the Crank-Nichilson scheme:

ϕn+1ϕiτ=H22(ϕn+1+ϕ),\frac{\phi_{n+1}-\phi^{\prime}}{-i\tau}=\frac{H_{2}}{2}\left(\phi_{n+1}+\phi^{\prime}\right), (23)

which gives

ϕn+1=1iτH2/21+iτH2/2ϕ.\phi_{n+1}=\frac{1-i\tau H_{2}/2}{1+i\tau H_{2}/2}\phi^{\prime}. (24)

Specifically, in the discretized coordinate space we have

H2ϕj=12m(ϕj+1+ϕj12ϕjδ2+1jδϕj+1ϕj12δ).H_{2}\phi_{j}=-\frac{1}{2m}\left(\frac{\phi_{j+1}+\phi_{j-1}-2\phi_{j}}{\delta^{2}}+\frac{1}{j\delta}\frac{\phi_{j+1}-\phi_{j-1}}{2\delta}\right). (25)

Here we have simplified ϕ(ρj)\phi(\rho_{j}) as ϕj\phi_{j}. To this end, (24) gives the updated wavefunction ϕn+1\phi_{n+1} at time tn+1t_{n+1}.

In above we have shown the details on the numerical simulation of GP equations. By transforming τiτ\tau\to-i\tau, i.e., the imaginary time evolution, one can also obtain the ground state of the system (droplet solution). To make sure that the wavefunction is reduced to zero well before touching the boundary of the system, we have chosen a sufficiently large size in the simulation with the maximum radius ranging from 50μm\mu m to 200μm\mu m. In this way one can avoid the boundary effect to dynamical and static properties of the system.

In solving the breathing modes from Eq.(14) in the main text, we have again discretized the coordinate space and transformed it to a large matrix equation. The breathing modes are then obtained by exactly diagonalizing the matrix.

References

  • [1] Y. S. Joung and C. R. Buie, Aerosol generation by raindrop impact on soil, Nat. Commun.6 , 6083 (2015).
  • [2] A. R. Vaezi, M. Ahmadi, and A. Cerdà, Contribution of raindrop impact to the change of soil physical properties and water erosion under semi-arid rainfalls, Sci. Total Environ. 583, 382 (2017).
  • [3] D. B. van Dam and C. Le Clerc, Experimental study of the impact of an ink-jet printed droplet on a solid substrate, Phys. Fluids 16, 3403 (2004).
  • [4] R. Mcpherson, The relationship between the mechanism of formation, microstructure and properties of plasma-sprayed coatings, Thin Solid Films 83, 297–310 (1981).
  • [5] M. R. O. Panão and A. L. N. Moreira, Intermittent spray cooling: a new technology for controlling surface temperature, Int. J. Heat Fluid Flow 30, 117–130 (2009).
  • [6] L. Mishchenko, B. Hatton, V. Bahadur, J. A. Taylor, T. Krupenkin, and J. Aizenberg, Design of ice-free nanostructured surfaces based on repulsion of impacting water droplets , ACS Nano 4, 7699 (2010).
  • [7] D. Khojasteh, M. Kazerooni, S. Salarian, and R. Kamali, Droplet impact on superhydrophobic surfaces: A review of recent developments, J. Ind. Eng. Chem. 42, 1 (2016).
  • [8] A. M. Worthington, XXVIII. On the forms assumed by drops of liquids falling vertically on a horizontal plate, Proc. R. Soc. Lond. 25, 261 (1877).
  • [9] A. M. Worthington, A study of splashes (Longmans, Green and Company, 1908).
  • [10] R. Rioboo, C. Tropea, and M. Marengo, Outcomes from a drop impact on solid surfaces, At. Sprays 11, 155 (2001).
  • [11] L. Xu, W. W. Zhang, and S. R. Nagel, Drop Splashing on a Dry Smooth Surface, Phys. Rev. Lett. 94, 184505 (2005).
  • [12] I. S. Bayer and C. M. Megaridis, Contact angle dynamics in droplets impacting on flat surfaces with different wetting characteristics, J. Fluid Mech. 558, 415–449 (2006).
  • [13] A. Latka, A. Strandburg-Peshkin, M. M. Driscoll, C. S. Stevens, and S. R. Nagel, Creation of Prompt and Thin-Sheet Splashing by Varying Surface Roughness or Increasing Air Pressure. Phys. Rev. Lett. 109, 054501 (2012).
  • [14] I.V. Roisman, A. Lembach, and C. Tropea, Drop splashing induced by target roughness and porosity: The size plays no role, Adv. Colloid Interface Sci. 222, 615 (2015).
  • [15] D. G. Aboud and A.-M.Kietzig, Splashing threshold of oblique droplet impacts on surfaces of various wettability. Langmuir 31, 10100 (2015).
  • [16] C. Tang, M. Qin, X. Weng, X. Zhang, P. Zhang, J. Li, and Z. Huang, Dynamics of droplet impact on solid surface with different roughness, Int. J. Multiph. Flow 96, 56 (2017).
  • [17] J. Hao, Effect of surface roughness on droplet splashing. Phys. Fluids 29, 122105 (2017).
  • [18] F. R. Smith, N. C. Buntsma, D. Brutin, Roughness influence on human blood drop spreading and splashing. Langmuir 34, 1143 (2017).
  • [19] T. C.de Goede, N. Laan, K. De Bruin, and D. Bonn, Effect of wetting on drop splashing of Newtonian fluids and blood. Langmuir 34, 5163 (2017).
  • [20] J. Tian and B. Chen, Dynamic behavior of non-evaporative droplet impact on a solid surface: Comparative study of R113, water, ethanol and acetone, Exp. Therm Fluid Sci. 105, 153 (2019).
  • [21] H. Almohammadi and A. Amirfazli, Droplet impact: Viscosity and wettability effects on splashing, J. Colloid Interface Sci. 553, 22 (2019).
  • [22] M.A. Quetzeri-Santiago, K. Yokoi, A.A. Castrejón-Pita, and J.R. Castrejón-Pita, Role of the Dynamic Contact Angle on Splashing, Phys. Rev. Lett. 122, 228001 (2019).
  • [23] H. Zhang, X. Zhang, X. Yi, F. He, F. Niu, and P. Hao, Effect of wettability on droplet impact: Spreading and splashing, Exp. Therm Fluid Sci. 124, 110369 (2021).
  • [24] L. Yang, Z. Li, T. Yang, Y. Chi, and P. Zhang, Experimental study on droplet splash and receding breakup on a smooth surface at atmospheric pressure, Langmuir 37, 10838 (2021).
  • [25] P. García-Geijo, E. Quintero, G. Riboux, and J. Gordillo, Spreading and splashing of drops impacting rough substrates J. Fluid Mech. 917 ,A50 (2021).
  • [26] A. Yarin, Drop impact dynamics: Splashing, spreading, receding, bouncing…, Annu. Rev. Fluid Mech. 38, 159 (2006).
  • [27] C. Josserand and S. Thoroddsen, Drop impact on a solid surface, Annu. Rev. Fluid Mech. 48, 365 (2016).
  • [28] I. Bloch, J. Dalibard and W. Zwerger, Many-body physics with ultracold gases, Rev. Mod. Phys. 80, 885 (2008).
  • [29] C. Chin, R. Grimm, P. Julienne and E. Tiesinga, Feshbach resonances in ultracold gases, Rev. Mod. Phys. 82, 1225 (2010).
  • [30] I. Ferrier-Barbut, H. Kadau, M. Schmitt, M. Wenzel, and T. Pfau, Observation of Quantum Droplets in a Strongly Dipolar Bose Gas, Phys. Rev. Lett. 116, 215301 (2016).
  • [31] M. Schmitt, M. Wenzel, F. Böttcher, I. Ferrier-Barbut, and T. Pfau, Self-bound droplets of a dilute magnetic quantum liquid, Nature 539, 259 (2016).
  • [32] I. Ferrier-Barbut, M. Schmitt, M. Wenzel, H. Kadau, and T. Pfau, Liquid quantum droplets of ultracold magnetic atoms, J. Phys. B 49, 214004 (2016).
  • [33] L. Chomaz, S. Baier, D. Petter, M.J. Mark, F. Wächtler, L. Santos, and F. Ferlaino, Quantum-Fluctuation-Driven Crossover from a Dilute Bose-Einstein Condensate to a Macrodroplet in a Dipolar Quantum Fluid, Phys. Rev. X 6, 041039 (2016).
  • [34] L. Tanzi, E. Lucioni, F. Fama, J. Catani, A. Fioretti, C. Gabbanini, R. N. Bisset, L. Santos, and G. Modugno, Observation of a Dipolar Quantum Gas with Metastable Supersolid Properties, Phys. Rev. Lett. 122, 130405 (2019).
  • [35] F. Böttcher, J.-N. Schmidt, M. Wenzel, J. Hertkorn, M. Guo, T. Langen, and T. Pfau, Transient Supersolid Properties in an Array of Dipolar Quantum Droplets, Phys. Rev. X 9, 011051 (2019).
  • [36] L. Chomaz, D. Petter, P. Ilzhöfer, G. Natale, A. Trautmann, C. Politi, G. Durastante, R.M.W. van Bijnen, A. Patscheider, M. Sohmen, M.J. Mark, and F. Ferlaino, Long-Lived and Transient Supersolid Behaviors in Dipolar Quantum Gases, Phys. Rev. X 9, 021012 (2019).
  • [37] C.R. Cabrera, L. Tanzi, J. Sanz, B. Naylor, P. Thomas, P. Cheiney, and L. Tarruell, Quantum liquid droplets in a mixture of Bose-Einstein condensates, Science 359, 301 (2018).
  • [38] P. Cheiney, C. R. Cabrera, J. Sanz, B. Naylor, L. Tanzi, L. Tarruell, Bright Soliton to Quantum Droplet Transition in a Mixture of Bose-Einstein Condensates, Phys. Rev. Lett. 120, 135301 (2018).
  • [39] G. Semeghini, G. Ferioli, L. Masi, C. Mazzinghi, L. Wolswijk, F. Minardi, M. Modugno, G. Modugno, M. Inguscio, M. Fattori, Self-Bound Quantum Droplets of Atomic Mixtures in Free Space, Phys. Rev. Lett. 120, 235301 (2018).
  • [40] C. D’Errico, A. Burchianti, M. Prevedelli, L. Salasnich, F. Ancilotto, M. Modugno, F. Minardi, and C. Fort, Observation of quantum droplets in a heteronuclear bosonic mixture, Phys. Rev. Res. 1, 033155 (2019).
  • [41] A. Burchianti, C. D’Errico, M. Prevedelli, L., Salasnich, F. Ancilotto, M. Modugno, F. Minardi, C. Fort, A dual-species Bose-Einstein condensate with attractive interspecies interactions, Condens. Matter, 5, 21 (2020).
  • [42] Z. Guo, F. Jia, L. Li, Y. Ma, J. M. Hutson, X. Cui, and D. Wang, Lee-Huang-Yang effects in the ultracold mixture of 23Na and 87Rb with attractive interspecies interactions, Phys. Rev. Res. 3, 033247 (2021).
  • [43] D. S. Petrov, Quantum Mechanical Stabilization of a Collapsing Bose-Bose Mixture, Phys. Rev. Lett. 115, 155302 (2015).
  • [44] D. S. Petrov and G. E. Astrakharchik, Ultradilute Low-Dimensional Liquids, Phys. Rev. Lett. 117, 100401 (2016).
  • [45] D. Edler, C. Mishra, F. Wächtler, R. Nath, S. Sinha, and L. Santos, Quantum Fluctuations in Quasi-One-Dimensional Dipolar Bose-Einstein Condensates, Phys. Rev. Lett. 119, 050403 (2017).
  • [46] K. Jachymski and R. Oldziejewski, Nonuniversal beyond mean field properties of quasi-two-dimensional dipolar Bose gases, Phys. Rev. A 98, 043601 (2018).
  • [47] P. Zin, M. Pylak, T. Wasak, M. Gajda, and Z. Idziaszek, Quantum Bose-Bose droplets at a dimensional crossover, Phys. Rev. A 98, 051603(R) (2018).
  • [48] T. Ilg, J. Kumlin, L. Santos, D. S. Petrov, and H. P. Büchler, Dimensional crossover for the beyond-mean-field correction in Bose gases, Phys. Rev. A 98, 051604(R) (2018).
  • [49] X. Cui, Y. Ma, Droplet under confinement: Competition and coexistence with a soliton bound state, Phys. Rev. Res. 3, L012027 (2021).
  • [50] X. Cui, Spin-orbit coupling induced quantum droplet in ultracold Bose-Fermi mixtures, Phys. Rev. A 98, 023630 (2018).
  • [51] S. Adhikari, A self-bound matter-wave boson–fermion quantum ball, Laser Phys. Lett 15, 095501 (2018).
  • [52] D. Rakshit, T. Karpiuk, M. Brewczyk, and M. Gajda, Quantum Bose-Fermi droplets, SciPost Phys. 6, 079 (2019).
  • [53] D. Rakshit, T. Karpiuk, P. Zin, M. Brewczyk, M. Lewenstein, and M. Gajda, Self-bound Bose-Fermi liquids in lower dimensions, New J. Phys. 21, 073027 (2019).
  • [54] M. Wenzel, T. Pfau and I. Ferrier-Barbut, A fermionic impurity in a dipolar quantum droplet, Phys. Scr. 93, 104004 (2018).
  • [55] J.-B. Wang, J.-S. Pan, X. Cui, W. Yi, Quantum droplet in a mixture of Bose-Fermi superfluids, Chin. Phys. Lett. 37, 076701 (2020).
  • [56] J. C. Smith, D. Baillie, and P. B. Blakie, Quantum Droplet States of a Binary Magnetic Gas, Phys. Rev. Lett. 126, 025302 (2021).
  • [57] R. N. Bisset, L. A. Pen~\tilde{n}a Ardila, and L. Santos, Quantum Droplets of Dipolar Mixtures, Phys. Rev. Lett. 126, 025301 (2021).
  • [58] Y. Ma, C. Peng, and X. Cui, Borromean Droplet in Three-Component Ultracold Bose Gases, Phys. Rev. Lett. 127, 043002 (2021).
  • [59] G. Ferioli, G. Semeghini, S. Terradas-Briansó, L. Masi, M. Fattori, and M. Modugno, Dynamical formation of quantum droplets in a 39K mixture, Phys. Rev. Res. 2, 013269 (2020).
  • [60] V. Cikojević ,L. Vranješ Markić, M. Pi , M. Barranco, F. Ancilotto, and J. Boronat, Dynamics of equilibration and collisions in ultradilute quantum droplets, Phys. Rev. Res. 3, 043139 (2021).
  • [61] C. Fort and M. Modugno, Self-evaporation dynamics of quantum droplets in a 41K-87Rb mixture, Appl. Sci. 11, 866 (2021).
  • [62] G. E. Astrakharchik and B. A. Malomed, Dynamics of one-dimensional quantum droplets, Phys. Rev. A 98, 013631 (2018).
  • [63] G. Ferioli, G. Semeghini, L. Masi, G. Giusti, G. Modugno,M. Inguscio, A. Gallemí, A. Recati, and M. Fattori, Collisions of Self-Bound Quantum Droplets, Phys. Rev. Lett. 122, 090401 (2019).
  • [64] R. Y. Chiao, E. Garmire, and C. H. Townes, Self-Trapping of Optical Beams, Phys. Rev.Lett. 13, 479 (1964).
  • [65] K. D. Moll, A. L. Gaeta, and G. Fibich, Self-Similar Optical Wave Collapse: Observation of the Townes Profile, Phys. Rev. Lett. 90, 203902 (2003).
  • [66] C. A. Chen and C. L. Hung, Observation of Universal Quench Dynamics and Townes Soliton Formation from Modulational Instability in Two-Dimensional Bose Gases, Phys. Rev. Lett. 125, 250401 (2020).
  • [67] C. A. Chen and C. L. Hung, Observation of Scale Invariance in Two-Dimensional Matter-Wave Townes Solitons, Phys. Rev. Lett. 127, 023604 (2021).
  • [68] B. Bakkali-Hassani, C. Maury, Y. Q. Zou, É. L. Cerf,R. Saint-Jalm, P. C. M. Castilho, S. Nascimbene, J. Dalibard, and J. Beugnon, Realization of a Townes Soliton in a Two-Component Planar Bose Gas, Phys. Rev. Lett. 127, 023603 (2021).
  • [69] Y.-Q. Zou, É. Le Cerf, B. Bakkali-Hassani, C. Maury, G. Chauveau, P.C.M. Castilho, R. Saint-Jalm, S. Nascimbene, J. Dalibard, and J. Beugnon, Optical control of the density and spin spatial profiles of a planar Bose gas, J. Phys. B: At. Mol. Opt. Phys., 54, 08LT01 (2021).
  • [70] C. K. Law, H. Pu, and N. P. Bigelow, Quantum Spins Mixing in Spinor Bose-Einstein Condensates, Phys. Rev. Lett. 81, 5257 (1998).
  • [71] X. L. Cui, Y. P. Wang and F. Zhou, Quantum-fluctuation-driven coherent spin dynamics in small condensates, Phys. Rev. A. 78, 050701(R) (2008).
  • [72] R. Barnett, J. D. Sau and S. Das Sarma, Antiferromagnetic spinor condensates are quantum rotors, Phys. Rev. A 82, 031602(R) (2010).
  • [73] J. Heinze, F. Deuretzbacher and D. Pfannkuche, Influence of the particle number on the spin dynamics of ultracold atoms, Phys. Rev. A 82, 023617 (2010).
  • [74] M. Koashi and M. Ueda, Exact Eigenstates and Magnetic Response of Spin-1 and Spin-2 Bose-Einstein Condensates, Phys. Rev. Lett. 84, 1066 (2000).
  • [75] T.-L. Ho and S. K. Yip, Fragmented and Single Condensate Ground States of Spin-1 Bose Gas, Phys. Rev. Lett. 84, 4031 (2000).
  • [76] E. J. Mueller, T.-L. Ho, M. Ueda, and G. Baym, Fragmentation of Bose-Einstein condensates, Phys. Rev. A 74, 033612 (2006).
  • [77] Q. Zhou and X. Cui, Fate of a Bose-Einstein Condensate in the Presence of Spin-Orbit Coupling, Phys. Rev. Lett. 110, 140407 (2013).