-
In terms of chiral fields, the Lagrangians of two massless noninteracting quarks u and d are invariant under the global
$S U(2)_L\times S U(2)_R$ chiral phase transformations$ \begin{array}{*{20}{l}} \psi _{L,R}\rightarrow \psi _{L,R}'=U_{L,R}\psi _{L,R}, \end{array} $
(1) where
$ \psi_{L,R}=\binom{u}{d}_{L,R} $ and$U_{L,R}=\exp (-{\rm i} \vec{\theta}_{L,R}\cdot \frac{\vec{\tau}}{2})$ . However, this chiral symmetry does not appear in the low energy particle spectrum, and the strong interaction theory exhibits the phenomenon of spontaneous symmetry breaking. Consequently, three Goldstone bosons appear, and the constituent quarks become massive at low energy. In describing the symmetries of the Lagrangian, it is useful to introduce three pion mesons$ \vec{\pi } $ and a σ meson in terms of a matrix field as$ \Phi =\sigma \frac{\tau ^0}2+ {\rm i} \vec{\pi }\cdot \frac{ \vec{\tau }}2, $
(2) where
$ \tau ^0 $ is the unity matrix and$ \vec{\tau } $ are the three Pauli matrices. Under the$S U(2)_L\times S U(2)_R$ chiral symmetry transformations, Φ transforms as$ \begin{array}{*{20}{l}} \Phi \rightarrow \Phi'= U_{L}\Phi U_{R}^+. \end{array} $
(3) Then, the renormalizable effective Lagrangian of the two-flavors quark meson model is defined as [9, 50]
$ \begin{array}{*{20}{l}} {\mathcal{L}}={\mathcal{L}}_\Phi +{\mathcal{L}}_q, \end{array} $
(4) where
$ {\mathcal{L}}_\Phi = \rm{Tr}[(\partial _\mu \Phi )^+(\partial ^\mu \Phi )]-\lambda \left[ {\rm Tr}(\Phi ^+\Phi )-\frac{{\vartheta}^{2} }{2}\right]^2-H \rm{Tr}[\Phi], $
(5) and
$ \begin{array}{*{20}{l}} {\mathcal{L}}_q=\overline{\psi }_L {\rm i} \not \partial \psi_L+\overline{\psi }_R {\rm i}\not \partial \psi _R-2g\overline{\psi }_L\Phi \psi _R+ {\rm h.c.}. \end{array} $
(6) Here, we have introduced a flavor-blind Yukawa coupling g of the left-handed and right-handed quark fields to interact with the Φ field.
The parameters of the Lagrangian
$ \mathcal{L} $ are chosen under the requirement that the chiral symmetry$S U(2)_L\times S U(2)_R$ is spontaneously broken down to$S U(2)_{L+R}$ in the vacuum, while the σ field takes on a non-vanishing vacuum expectation value$\langle\sigma\rangle ={\it f}_{\pi}=93 ~\mathrm{MeV}$ . This results in a massive σ meson and three massless Goldstone bosons$ \vec{\pi } $ mesons in the chiral limit, as well as giving an effective mass$ m_q=gf_\pi $ to the constituent quarks. Furthermore, the chiral symmetry is explicitly broken by adding the last term in Eq. (5) due to the finite current quark masses. With this additional term, the vector isospin$S U(2)$ symmetry remains exact but the axial$S U(2)$ transformation is no longer invariant. Accordingly, the constant H is to be fixed by the partially conserved axial vector current relation, which gives$ H=f_{\pi}m_{\pi}^{2} $ , where the pion mass is taken as$ m_{\pi}=138 $ MeV. Moreover, the dimensionless coupling constant g in the model is determined by the constituent quark mass in vacuum, which is about$ 1/3 $ of the nucleon mass and gives$ g\simeq 3.3 $ . Another dimensionless coupling constant λ is usually fixed by the sigma mass$ m_\sigma^2=m^2_\pi+2\lambda f^2_\pi $ . Here, we set it to$ 500 $ MeV according to the most recent compilation of the Particle Data Group [51]. Finally, the quantity ϑ is actually not a free parameter and can be formally expressed as$ \vartheta^{2}=f^{2}_{\pi}-m^{2}_{\pi}/\lambda $ . -
A convenient framework for studying phase transitions and the restoration of the chiral symmetry at extremely high energy is the thermal field theory [18, 52]. Within this framework, the effective potential is one of the important and powerful theoretical tools, and the standard approach for dealing with the thermodynamics of various observables of interest relies on the grand canonical ensemble. To make things lucid, we start with a spatially uniform system in thermodynamical equilibrium at temperature T and quark chemical potential μ. From here and henceforward, we will use the chemical potential to represent the quark chemical potential. In general, the grand partition function is commonly given by
$ \begin{array}{*{20}{l}} \mathcal{Z}= \displaystyle\int\prod_a \mathcal{D} \sigma \mathcal{D} \pi_a \displaystyle\int \mathcal{D}\psi \mathcal{D} \bar{\psi} \mathrm{exp} \left[ \int_X (\mathcal{L}+\mu \bar{\psi} \gamma^0 \psi ) \right], \end{array} $
(7) where
$\int_X\equiv \int^{\beta}_0 {\rm d}\tau \int {\rm d}^3x$ , the inverse temperature$ \beta=1/T $ , and$ \mu=\mu_B /3 $ for the homogeneous background field.In the mean-field approximation, the meson fields in the Lagrangian are replaced by their expectation values, whereas the quark and antiquark fields are still retained as quantum fields. This implies that the one-loop correction to the effective potential from the quark fields is considered but treats the mesonic degrees of freedom at the tree level. Following this scheme, the integration over the fermions yields a determinant that can be calculated using standard procedures [27, 53], generating an effective potential for the mesons. Finally, the effective potential of the model can be obtained exactly in a closed form as
$ \Omega(T,\mu)=\frac{-T \mathrm{ln} \mathcal{Z}}{V}=U(\sigma ,\vec{\pi} )+\Omega_{\bar{\psi} \psi}, $
(8) where the classical potential for the σ and
$ \vec{\pi} $ is rewritten as$ U(\sigma ,\vec{\pi})=\frac{\lambda}{4} \left(\sigma ^{2}+\vec{\pi} ^{2} -{\vartheta}^{2}\right)^{2}-H\sigma, $
(9) and the contribution of quarks and antiquarks are given by
$ \begin{aligned}[b] \Omega_{\bar{\psi} \psi} =\;& \Omega_{\bar{\psi} \psi}^\mathrm{v}+\Omega_{\bar{\psi} \psi}^{th} = -\nu \int \frac{{\rm d}^3\vec{p}}{(2 \pi)^3} E \\& -\nu T \int \frac{{\rm d}^3\vec{p}}{(2 \pi)^3} \left\{ \mathrm{ln} \left[ 1+ {\rm e}^{-(E-\mu)/T}\right] +\mathrm{ln} \left[ 1+ {\rm e}^{-(E+\mu)/T}\right]\right\}. \end{aligned} $
(10) Here,
$ \nu=2N_f N_c=12 $ and$ E=\sqrt{\vec{p}^2+m_q^2} $ is the valence quark and antiquark energy for u and d quarks, and the minus sign is the consequence of the Fermi-Dirac statistics. The constituent quark (antiquark) mass is set to$ m_q=g\sigma $ .The first term of Eq. (10) denotes the fermions vacuum one-loop contribution, which is ultraviolet divergent and can only be evaluated in the presence of a regulator. The divergence in Eq. (10) can then be appropriately renormalized using the dimensional regularization scheme [47, 48, 54]. After considering the vacuum fluctuations and renormalization issues, the renormalized fermion vacuum one-loop contribution reads
$ \Omega_{\bar{\psi} \psi} ^{\mathrm{v}}=\Omega_{\bar{\psi} \psi} ^{\mathrm{reg}}=-\frac{N_c N_f}{8\pi^2} m_q^4 \mathrm{ln}\left(\frac{m_q}{\Lambda}\right), $
(11) where Λ denotes the arbitrary renormalization scale. Notably, dimensional regularization introduces an arbitrary renormalization scale parameter. Nevertheless, at least in the one-loop approximation, the thermodynamic potential and all physical observables are independent of the choice of Λ, and the scale dependence can be neatly canceled out after the rearrangement of parameters in the model [40, 47, 48, 55].
Equipped with the above effective potential, we can explore the phase diagram of the model at finite temperature and density by minimizing the thermodynamical potential in equation (8) with respect to the order parameter σ. Then, an equation of motion is given by
$ \frac{\partial\Omega(T,\mu)}{\partial \sigma}=0. $
(12) The solution of the equation of motion determines the behavior of the chiral order parameters σ as a function of T and μ, as well as the phase diagram of the model. As we know, the thermodynamic state of equilibrium is determined by the values of the order parameter at the global minimum of the effective potential. Once the order parameter for each given T and μ is obtained, any thermodynamical quantity of equilibrium, such as the pressure, the entropy density, the energy density, the speed of sound, et al., can be described and calculated.
In Fig. 1, we have presented the phase diagram in calculation with the fermion vacuum fluctuation for the two-flavor quark meson model. The temperature behavior of the chiral condensate σ shows that the system experiences a smooth crossover transition at low chemical potential, while there is a first-order phase transition for larger chemical potential because the chiral order parameter makes a jump across the gap of the condensate near the critical temperature
$ T_c $ . Normally, the temperature derivative of the chiral condensate σ for quarks has a peak at some specific temperature, which is established as the critical temperature for the chiral phase transition. Because the temperature derivative of the chiral condensate has simply one peak, we can not tell when and where the crossover phase transition would convert to a first-order one at the critical endpoint (CEP) with a second order phase transition [46, 50]. In order to locate the CEP in the phase diagram, the quark number susceptibility$\chi_q= \partial^2 \Omega(T,\mu)/ \partial^2 \mu$ is to be introduced, and it is believed to be divergent at the CEP [4, 5].Figure 1. (color online) The phase diagram in the
$ T-\mu $ plane for the two-flavor quark meson model. The dashed lines are the critical line for conventional chiral phase transition in the crossover region. The solid line indicates the first-order phase transitions, and the solid circle indicates the critical end points for chiral phase transitions of u and d quarks. The dashed-dotted line and the dashed-doted-dotted line are the lower and upper spinodal lines.Aside from calculation of the quark number susceptibility
$ \chi_q $ , in the present work, we prefer to use the shapes of the effective potential at various temperatures and chemical potentials to decide the position of the CEP. In the case of the first-order phase transition, along the critical line with the temperature$ T\simeq T_c $ , the thermodynamical potential$ \Omega(T,\mu) $ has two minima of equal depth separated by a potential barrier. With the reduction of the chemical potential, the height of the barrier decreases and finally disappears at the CEP, where the phase transition is of the second order. In our calculation, the corresponding CEP is located at$ (T_E, \mu_E)\simeq (30,301) $ MeV in Fig. 1. Notably, the location of the CEP from the theory calculations is scattered over the region of$ \mu_B=200-1100 $ MeV and$ T=40-180 $ MeV [4, 5]. QCD-based model calculations like the NJL model [56, 57], QM model [58, 59], PNJ [10], and PQM [12−14] produce a relative larger critical chemical potential around$ \mu=\mu_B/3=300 $ MeV. However, the functional renormalization group (FRG) approach and Dyson-Schwinger equations predict a rather narrow region for the critical chemical potential around$\mu=200-220$ MeV [60−63]. However, the accuracy of predictions for CEP from the first principle lattice-QCD calculations worsens toward a very large chemical potential; various model calculations vary wildly in their predictions. Therefore, an experimental search of the critical point is crucial and important to establish its position in the phase diagram.As shown in Fig. 2, in the region of the first-order phase transition, a typical effective potential commonly displays a local minimum at a low sigma
$ \sigma_l $ , which is separated by a potential barrier from another local minimum at a relative larger sigma$ \sigma_h $ . When a critical temperature$ T_c $ is reached, these two minima degenerate. For$ T<T_c $ , the minimum of the effective potential at$ \sigma=\sigma_h $ is the absolute or global minimum, which is regarded as the stable (true) vacuum, whereas the minimum at$ \sigma=\sigma_l $ is treated as the metastable (false) vacuum. In this case, the chiral symmetry is broken so that the constituent quarks become massive. On the contrary, when the temperature T goes across above the critical value$ T_c $ , these two vacua flip over, the global minimum is at$ \sigma=\sigma_l $ , and the local minimum is at$ \sigma=\sigma_h $ . Since the chiral symmetry is approximately restored and the quarks become almost massless, the system for$ T>T_c $ is then considered as the quark phase. The previous case for$ T<T_c $ is taken as the hadron phase, therefore the critical lines divide the whole phase diagram into two categories: the hadron and quark phases.Figure 2. (color online) (a) The grand canonical potentials Ω as a function of the chiral order parameter σ for
$ \mu=306 $ MeV at various temperatures. (b) The grand canonical potentials Ω as a function of the chiral order parameter σ for$ \mu=309 $ MeV at various temperatures.Normally, apart from the critical temperature
$ T_c $ , there are two other temperatures of interests in a first-order phase transitions. These two temperatures$ T_{c1} $ and$ T_{c2} $ are named the lower and upper spinodal critical points, respectively. A typical example is shown in Fig. 2, where the evolutions of the potential for several temperatures when a chemical potential fixed at$ \mu=306 $ MeV and$ \mu=309 $ MeV are exhibited. For the left panel in Fig. 2 at$ \mu=306 $ MeV, when the temperature is around$ T_c\simeq20.6 $ MeV, the shape of the potential exhibits two degenerate minima. However, as the temperature increases, the second minimum of the potential at$ \sigma=\sigma_h $ disappears at a higher temperature$ T_{c2}\simeq 23.1 $ MeV. Meanwhile, when the temperature falls below the critical temperature$ T_c $ , the first minimum of the potential at$ \sigma=\sigma_l $ tends to wipe out around$ T_{c1}\simeq 14.7 $ MeV. Between these two specific temperatures, metastable states or a false vacuum exists, and the system can exhibit supercooling or superheating.For
$ \mu=309~ \mathrm{MeV} $ , one can also observe the characteristic pattern of a first order phase transition: two minima corresponding to phases of restored and broken chiral symmetry are separated by a potential barrier and they become degenerate when the temperature is at$ T_c \simeq 13.3 $ MeV. Chiral symmetry is approximately restored for$ T>T_c $ , where the minimum at false vacuum$ \sigma=\sigma_l $ becomes the absolute minimum, as shown in the right panel in Fig. 2. Similar to the previous case for$ \mu=306 $ MeV, when the temperature T goes up the critical line and rises further, the potential barrier between two minima decreases gradually and shrinks to zero at the moment when the second minimum of the potential at$ \sigma=\sigma_h $ vanishes at a spinodal temperature$ T_{c2}\simeq 18.8 $ MeV. On the other hand, when$ T<T_c $ , the shapes of the effective potential at various low temperatures display quite different behaviors in comparison with those of the previous case at$ \mu=306 $ MeV. The barrier between two minima of the effective potential is to maintain even when the temperature T is very close to zero. This means that the first minimum of the effective potential at$ \sigma=\sigma_l $ could always exist in the hadron phase. Therefore, the phase transition could be identified as a strongly first-order phase transition, which is usually induced by an effective potential with a nonvanishing zero-temperature potential barrier.To provide a complete description of a first-order phase transition, two particular lines of the spinodal points that constrain the regions of spinodal instability for the first-order phase transition at high density are illustrated in Fig. 1. Similar to the critical line in Fig. 1, both the lower and upper spinodal lines increase as the chemical potential μ reduces. However, the gap between these two spinodal lines becomes increasingly smaller. In the end, the two spinodal lines and the critical line terminate at the same point in the CEP. Moreover, since the lower spinodal line will end up at some point
$ \mu_c\simeq 308 $ MeV on the vertical axis of the chemical potential, the area of the first-order phase transition can be technically split into a weakly first-order phase transition and a strongly first-order phase transition according to the above discussion.Therefore, for a weak first-order chiral phase transition where the chemical potential is
$ \mu<\mu_c $ at$ T<T_c $ , the thermodynamic potential exhibits a local minimum aside from the global minimum, when the temperature decreases from$ T_c $ to a specific value$ T_{c1} $ . The local minimum gradually disappears at the point of inflection known as spinodal instability. Whereas, for$ \mu>\mu_c $ , the chiral phase transition is to be considered as a strongly first-order one due to the fact that the local minimum remains for the temperature at$ T<T_c $ and there is no spinodal temperature. The critical chemical potential for a transition from a weak first-order phase transition to that of a strong one is then identified as the critical chemical potential at$ \mu_c\simeq 308 $ MeV in hadron phase [37, 64]. -
The mechanism of the nucleation theory can be used to study the probability that a bubble or droplet of the stable vacuum in a system initially trapped in the metastable vacuum near the critical temperature
$ T_c $ . For a pure system, the formation of bubbles originates from intrinsic thermodynamic fluctuations; this kind of nucleation mechanism is commonly called homogeneous nucleation. On the contrary, when impurities cause the formation of bubbles or droplets, such a mechanism of the nucleation theory is known as heterogeneous nucleation. In the everyday world, the external agents would play the role of nucleating centers, such as dust or ions in the atmosphere, leading to a much more efficient increase in the nucleation rate. Nevertheless, for the physical interests related to our study, homogeneous nucleation theory is appropriate, and we will use this basic theoretical apparatus to describe the decay of the metastable vacuum of a system interacting with a heat bath at temperature T.Based on the framework of the homogeneous thermal nucleation, we make an assumption in the limit that thermal fluctuations dominate quantum fluctuations, and the quantum-induced tunneling is simply ignored. Then, the nucleation rate per unit time per unit volume is given in the form of
$ \Gamma=\mathcal{P}\exp\left[ -\frac{S_3}{T} \right], $
(13) where T is the temperature of the system in equilibrium with the thermal bath,
$ S_3 $ is the three-dimensional action associated with the$ O(3) $ -symmetric critical bubble or droplet and$ \mathcal{P} $ is the exponential prefactor. For the mechanism of the bubble nucleation to the leading order, the nucleation rate is controlled by the exponent of the three-dimensional action evaluated on the critical bubble. The sub-leading corrections to the leading-order bubble action are included in the prefactor$ \mathcal{P} $ , which can be technically expressed as$ \mathcal{P}= \frac{\omega_-}{2\pi}\left( \frac{S_3}{2\pi T} \right)^{3/2}\left[ \frac{\mathrm{Det}(-\nabla^2+\Omega''_{FV})}{\mathrm{Det}'(-\nabla^2+\Omega''_{B})} \right]^{1/2}. $
(14) Here,
$ \omega_- $ is the eigenvalue of the negative mode, the terms$ \Omega''_{FV} $ and$ \Omega''_{B} $ are abbreviations for$ \Omega'' $ evaluated in the false vacuum and the critical bubble, the prime in the determinant signifies that the zero eigenvalues associated with the translation symmetry of the bubble are omitted.$ \Omega'' $ is the second derivative of the effective potential$ \Omega(T,\mu) $ with respect to the order parameter σ which is actually represented a field in describing the extremum of the three-dimensional Euclidean action, more specifically, the critical bubble or the bounce. Usually, the calculation and evaluation of this prefactor is a nontrivial matter, a rough estimate of their ratios can be obtained by dimensional analysis and it can be approximately expressed as$ T^4 $ or$ T^4_c $ for simplicity [26, 41].The result represented in the above equation (13) is a semi-classical contribution based on a saddle point approximation around the bounce solution. By taking the scalar field σ as the order parameter, at finite-temperature field theory, an Euclidean action we are interested in is
$ S_E(\sigma)=\int_0^{\beta} {\rm d} \tau \int {\rm d}^3r \left[ \frac{1}{2} \left(\frac{\partial \sigma}{\partial \tau} \right)^2+\frac{1}{2} \left(\nabla \sigma \right)^2+\Omega(\sigma;T,\mu) \right], $
(15) in which the subscript E denotes Euclidean and the integral is over Euclidean space. For the sake of convenience, in the following discussions, we will keep the σ field in the effective potential Ω in Eq. (8) explicitly. As argued by Linde [35], for sufficiently high temperatures at large length scales compared to β, the relevant number of dimensions is
$ d=3 $ , and the Euclidean action becomes$ S_E(\sigma)\equiv\frac{S_3}{T}, $
(16) where
$ S_3 $ is the three-dimensional saddle-point action associated with the formation of a critical-sized bubble or droplet; in what follows, it is to be called as the saddle-point action for abbreviation. Therefore, the bounce is an$ O(3) $ symmetric solution to the classical equation of motion that extremizes the Euclidean action$ S_3 $ . In particular, for a scalar field σ, the bounce satisfies a nonlinear ordinary differential equation,$ \frac{{\rm d}^2\sigma(r)}{{\rm d}r^2}+\frac{2}{r}\frac{{\rm d}\sigma(r)}{{\rm d}r}=\frac{\partial\Omega(\sigma;T,\mu)}{\partial \sigma}, $
(17) with boundary conditions
$ \lim\limits_{r \to \infty }\sigma(r)=\sigma_{FV} $ and$\dfrac{{\rm d}\sigma(r)}{{\rm d} r}\Big|_{r=0}= 0$ . The first boundary condition is, because the bubbles are embedded in the homogeneous false vacuum outside the bubble, the σ field should arrive at its false vacuum at$ \sigma \simeq \sigma_{FV} $ . The second one is set according to the requirement of finite energy at the origin. The solution for this equation of motion with the above proper boundary conditions is a saddle point solution or a bounce$ \sigma_b $ . It is an$ O(3) $ non-trivial field configuration that starts in the false vacuum and reaches the other side of the potential barrier with zero velocity. In this work, we will use the AnyBubble package [65] to determine the bounce.Once the solution
$ \sigma_b $ is obtained, the$ S_3 $ exponent in Eq. (13) can be evaluated on the bounce solution$ \sigma_b $ as$ S_3=\int {\rm d}^3r \left[ \frac{1}{2} \left(\nabla \sigma \right)^2+\Omega(\sigma;T,\mu) \right], $
(18) and the surface tension of the nucleation bubble interface between the false vacuum and the true vacuum is defined accordingly as [35, 66]
$ \Sigma=\int {\rm d} r \left[ \frac{1}{2} \left(\frac{{\rm d}\sigma}{{\rm d}r} \right)^2+\Omega(\sigma;T,\mu) \right]. $
(19) Notably, in practical calculations, if the false vacuum has a non-zero potential energy, an additional term
$ -\Omega(\sigma_{FV};T,\mu) $ should be included in the$ S_3 $ action and the surface tension Σ.For a generic effective potential, the equation of motion of the bounce with boundary conditions usually cannot be obtained analytically. We should rely on numerical methods to perform the computation. However, when the system is very close to the critical coexistence line, the bubble radius R is much larger than the wall thickness (
$ \triangle R \sim m_{\sigma}^{-1} $ ). Hence, when the damping force$ 2 \sigma'/r $ in the field equation becomes negligible, the thin-wall approximation is applicable and the equation of motion (17) reduces to the field equation for a typical one-dimensional soliton$ \frac{{\rm d}^2 \sigma (r)}{{\rm d}r^2}=\frac{{\rm d}\Omega}{{\rm d} \sigma}. $
(20) This static field equation implies that
$ \frac{{\rm d} \sigma(r)}{{\rm d}r}=\pm \sqrt{2 \Omega}. $
(21) Integrating Eq. (21) yields
$ r = \int_{\sigma}^{\sigma_{Fv}} \frac{{\rm d} \sigma}{\sqrt{2 \Omega}}. $
(22) Therefore, an approximate solution for the bounce can be obtained for arbitrary potential with two or more degenerate minima. Moreover, within the thin-wall approximation, the surface tension of the bubble can be calculated as
$ \Sigma_{\mathrm{tw}}=\int_{0}^{\infty}{\rm d}r \left[ \frac{1}{2} \left(\frac{{\rm d}\sigma}{{\rm d}r} \right)^2+\Omega \right]=\int_{\sigma_T}^{\sigma_{Fv}} {\rm d}\sigma \sqrt{2 \Omega}, $
(23) and the saddle-point action
$ S_3 $ is given by$ S_3=\frac{16 \pi}{3} \frac{\Sigma_{\mathrm{tw}}^3}{\varepsilon}. $
(24) The quantity
$\varepsilon=\Omega({\sigma_{Fv}};T,\mu)-\Omega(\sigma_T;T,\mu)$ is the difference between the values of the effective potential at the false vacuum and true vacuum.
Bubble nucleation in the two-flavor quark-meson model
- Received Date: 2023-11-27
- Available Online: 2024-05-15
Abstract: We investigate the dynamics of a first-order quark-hadron transition via homogeneous thermal nucleation in the two-flavor quark-meson model. The contribution of the fermionic vacuum loop in the effective thermodynamics potential and phase diagram, together with the location of the critical endpoint (CEP), is obtained in the temperature and chemical potential plane. For weak and strong first-order phase transitions, by taking the temperature as a variable, the critical bubble profiles, evolutions of the surface tension, and saddle-point action in the presence of a nucleation bubble are numerically calculated in detail when fixing the chemical potentials at