-
The evolution of the wave function
$ \psi({ r},t) $ of$ \Upsilon(1S) $ at time t and the relative radius$ { r} $ between the bottom quark and the anti-bottom quark can be described by the Schroedinger equation$ {\rm i}\partial_t \psi({ r},t) = \left[-\frac{1}{m_b}\nabla^2+V(r,T(t))-{\rm i}\Gamma({ r}, T)\right]\psi({ r},t), $
(1) where
$ m_b $ is the mass of a bottom quark, and we take$ \hbar = 1 $ . In the above, thermal fluctuation [33] is neglected. The corresponding stationary radial Schroedinger equation for a given temperature T is$ \left[-\frac{1}{m_br}\frac{ \rm{d}^2}{ \rm{d} r^2}r+\frac{l(l+1)}{m_br^2}+V(r,T)-{\rm i}\Gamma(r,T)\right]\psi_r(r) = E\psi_r(r),$
(2) where
$ \psi_r(r) $ is the radial wave function of$ \Upsilon(1S) $ , and E is the eigen energy of$ \Upsilon(1S) $ . For$ \Upsilon(1S) $ , the azimuthal quantum number is$ l = 0 $ . To focus on the novel mechanism, we neglect the particle scattering process and consider that the in-medium width$ \Gamma = 0 $ . The potential V is taken in the form of a screened Cornell potential [34]$ V(r)= -\frac{\alpha}{r}{\rm e}^{-\mu r}-\frac{\sigma}{2^{3/4}\Gamma(3/4)}\left(\frac{r}{\mu}\right)^{1/2}K_{1/4}[(\mu r)^2], $
(3) where
$ \alpha = \frac{\pi}{12} $ and$ \sigma = 0.2 $ GeV$ ^2 $ [34]. Further,$ \Gamma $ and K are the gamma function and the modified Bessel function, respectively. For simplicity, we neglect the constant term that does not vanish at infinity in$ V(r) $ . The dependence on temperature T arises from the screening mass$ \mu $ . We fit the free energy of heavy quarks by the lattice QCD [34, 35], and parameterize the screening mass$ \mu $ (scaled by$ \sqrt{\sigma} $ ) as$ \frac{\mu(\bar{T})}{\sqrt{\sigma}} = s\bar{T}+a\sigma_t\sqrt{\frac{\pi}{2}}\left[ {\rm erf}\left(\frac{b}{\sqrt{2}\sigma_t}\right)- {\rm erf}\left(\frac{b-\bar{T}}{\sqrt{2}\sigma_t}\right)\right], $
(4) with
$ \bar{T} = T/T_c $ ,$ s = 0.587 $ ,$ a = 2.150 $ ,$ b = 1.054 $ ,$ \sigma_t = $ 0.07379, and the error function${\rm erf}(z) = \frac{2}{\sqrt{\pi}}\int_0^{z} {\rm e}^{-x^2} {\rm d} x$ . Here,$ T_c $ is the critical temperature of the phase transition.The radial eigen wave function is shown in Fig. 1. To be dimensionless, we scaled the radius and the wave function as
$ \bar{r} = m_b r $ and$ \bar{\psi}_r = m_b^{-3/2}\psi_r $ , respectively, resulting in$\int \left|\bar{\psi}_r\right|^2 \bar{r}^2 {\rm d} \bar{r} = 1$ . The wave function of$ \Upsilon(1S) $ at$ T = T_c $ is similar to that at$ T = 0 $ , while it becomes increasingly broad with increasing temperature. The dissociation temperature is$ T_d\approx3T_c $ .Figure 1. (color online) Scaled radial wave functions
$\bar{\psi}_r = m_b^{-\frac{3}{2}}\psi$ of$ \Upsilon(1S) $ as a function of scaled radius$ \bar{r} = m_b r $ , at different scaled temperature$ \bar{T} = T/T_c $ .The transition probability from a
$ \Upsilon(1S) $ at zero temperature to that at T is$ P(\bar{T}) = \left|\left\langle \psi(T)|\psi(0)\right\rangle\right|^2,$
(5) which is shown as a function of
$ \bar{T} = T/T_c $ in Fig. 2. It decreases monotonically with$ \bar{T} $ , as the overlap between the wave function at finite temperature and that at zero temperature becomes small when$ \bar{T} $ increases. It is very close to unity at$ \bar{T} = 1 $ , as already indicated by Fig. 1, and it vanishes at$ T_d\approx3T_c $ . Note that excited states of$ \Upsilon(2S) $ and$ \Upsilon(3S) $ at temperature T can also be generated from$ \Upsilon(1S) $ in vacuum if T is below their dissociation temperatures$ T_d $ , and they can finally feed down to$ \Upsilon(1S) $ . This effect is not included in our current calculation.Figure 2. Transition probability P [defined in Eq. (5)] of a
$ \Upsilon(1S) $ from temperature$ 0 $ to temperature T as a function of scaled temperature$ \bar{T} = T/T_c $ .Subsequently, we verify the adiabatic approximation. At RHIC energy, the highest temperature of the fireball is around
$ 2T_c $ when the system reaches local thermal equilibrium. We suppose that the temperature decreases with time linearly from$ 2T_c $ to$ T_c $ , and evolve the wave function$ \psi({ r},t) $ of a$ \Upsilon(1S) $ by Eq. (1) with its initial condition as an eigenstate$ \Upsilon(1S) $ at$ 2T_c $ . The survival probability as a function of time is shown in Fig. 3. The typical time for the fireball to cool down to$ T_c $ is$ 5\sim10 $ fm/c. As shown in the figure, the survival probability is approximately$ 0.98 $ when the evolution time is$ 10 $ fm/c, which implies that the adiabatic approximation is very good in this case. Even if we take a lower value of$ 5 $ fm/c, the survival probability$ 0.93 $ is obviously larger than$ P(2.0) = 0.76 $ , as shown in Fig. 2. This result is qualitatively consistent with the result in Ref. [36], where the adiabatic approximation is examined for$ \Upsilon(1S) $ at LHC energy with a finite dissociation rate.Figure 3. (color online) Upper panel: Different cooling systems with medium temperature decreasing linearly with time. Lower panel: Time evolution of survival probability of
$ \Upsilon(1S) $ at different cooling speeds calculated using Schroedinger equation with an initial$ \Upsilon(1S) $ at its eigenstate at the initial temperature.Here, we include the spatial distribution of temperature. In practice, the temperature is not uniform in space. The temperature is high in the center of the fireball, whereas it is low in peripheral regions. Therefore, the survival probability for
$ \Upsilon(1S) $ is an average of all generated$ \Upsilon(1S) $ s. Because the production of$ \Upsilon(1S) $ is a hard process, we assume that the density of generated$ \Upsilon(1S) $ is proportional to the number density of binary collisions$ n_c({ x}_T) $ at transverse coordinate$ { x}_T $ . Therefore, we have$R_{AA} = \frac{\displaystyle\int P(\bar{T}({ x}_T)) {\rm d} N_{\Upsilon(1S)}}{\displaystyle\int {\rm d} N_{\Upsilon(1S)}} = \frac{\displaystyle\int P(\bar{T}({ x}_T))n_c({ x}_T) {\rm d} { x}_T}{\displaystyle\int n_c({ x}_T) {\rm d} { x}_T}. $
(6) We assume that the entropy density s is proportional to the density of the number of participants
$ n_p $ , and we regard the hot medium as an ideal gas, such that the entropy density is likewise proportional to$ T^3 $ . Consequently, the spatial distribution of temperature is$ \bar{T}({ x}_T)= \bar{T}({\bf 0})\left(\frac{n_p({ x}_T)}{n_p({\bf 0})}\right)^{1/3}, $
(7) where
$ \bar{T}({ x}_T) $ is the scaled local temperature$ T/T_c $ at$ { x}_T $ , and$ \bar{T}({\bf 0}) $ is the scaled local temperature at$ { x}_T = {\bf 0} $ . In central collisions, the number density of participants$ n_p $ and number density of binary collisions$ n_c $ are$ n_p({ x}_T) =2{\cal T}(x_T)\left[1-{\rm e}^{-\sigma_{NN}{\cal T}(x_T)}\right], $
(8) $ n_c({ x}_T) = \sigma_{NN}{\cal T}^2(x_T), $
(9) where
$ \sigma_{NN} $ is the inelastic cross-section of nucleons, and$ {\cal T}(x_T) $ is the thickness function of a gold nucleus. For simplicity, we take a sharp-cut-off thickness function$ {\cal T}(x_T)= \frac{3A\sqrt{R^2-x_T^2}}{2\pi R^3}, $
(10) where R and A are the radius and mass number of the nucleus, respectively. By substituting Eqs. (7)-(10) to Eq. (6), we obtain the nuclear modification factor in central collisions
$ R_{AA} = 4\int_0^1P\left(\bar{T}(0)\sqrt[\root{3}3]{x\frac{1-{\rm e}^{-N_mx}}{1-{\rm e}^{-N_m}}}\right)x^3 {\rm d} x, $
(11) with
$ N_m = \sigma_{NN}{\cal T}(x_T = 0) = 3\sigma_{NN}A/(2\pi R^2) $ , and$ \begin{array}{l} P(\bar{T}) = \left\{\begin{array}{ll}0,&\bar{T}>T_d/T_c,\\ \left|\langle \psi(T)|\psi(0)\rangle\right|^2,&1<\bar{T}<T_d/T_c,\\ 1,&\bar{T}<1.\end{array}\right. \end{array} $
(12) where we take
$ P = 1 $ below$ T_c $ as an approximation. We take$ R = 6.38 $ fm and$ A = 197 $ for gold [37], and$ \sigma_{NN} = 41 $ mb at RHIC energy [25].$ R_{AA} $ as a function of$ \bar{T}({\bf 0}) $ is shown in Fig. 4. The$ R_{AA} $ is above$ 0.9 $ if the central temperature$ T({\bf 0}) $ is lower than$ 1.6T_c $ , while it is below$ 0.8 $ when$ T({\bf 0}) $ is higher than$ 2.1T_c $ . We can expect that this effect is not negligible at the RHIC, and that it is considerable at the LHC. The factor$ x^3 $ in Eq. (11) arises from two facts: 1) more$ \Upsilon(1S) $ s are generated at the center of the fireball, and 2) the thickness changes slowly with the radius at the center of the fireball. Consequently,$ R_{AA} $ relies more on the survival probability P at the center of the fireball, i.e., at$ x_T = 0 $ . Therefore, the qualitative behavior of$ R_{AA} $ in Fig. 4 is similar to P in Fig. 2, and they are quantitatively similar when$ P(\bar{T}({\bf 0})) $ is large. From the hydrodynamics simulation, the initial maximum temperature is approximately$ 350 $ MeV at proper time$ \tau_0 = 0.6 $ fm/c, and it becomes$ 250 $ MeV at$ \tau = 2 $ fm/c, while the critical temperature in the same calculation is approximately$ 164 $ MeV [38]. We estimate that$ T({\bf 0}) $ is between these two values, and the corresponding$ \bar{T}({\bf 0}) $ is between$ 1.5 $ and$ 2.1 $ .Figure 4. Nuclear modification factor
$ R_{AA} $ in central Au+Au collisions due to heating dissociation effect as a function of scaled temperature$ \bar{T}({\bf 0}) $ at center of fireball.We provide three remarks concerning this result. 1) Even if the width (or dissociation rate)
$ \Gamma $ vanishes at finite temperature, there is a fast heating dissociation effect for$ \Upsilon(1S) $ suppression, which was not carefully considered before. 2) If the width of$ \Upsilon(1S) $ is negligible as in some calculations, then the heating dissociation of$ \Upsilon(1S) $ can be used as a thermometer to detect the temperature of the fireball at an early time, and it is not sensitive to the temperature later on. It is necessary to clarify that such a temperature measured via$ \Upsilon(1S) $ should never be interpreted as the highest temperature of the fireball, but rather as the temperature felt by a$ \Upsilon(1S) $ . As a matter of fact, the highest temperature at a very early time is not well defined, and the change in the temperature at this time occurs so rapidly that the adiabatic theorem becomes invalid [39], which means that$ \Upsilon(1S) $ may not feel the temperature before it drops relatively slowly. Indeed, the most interesting temperature is not the high and short-lived temperature at the very beginning, but the temperature that can be felt by particles. In this sense, the$ \Upsilon(1S) $ -felt temperature of the medium is more meaningful. 3) The transverse momentum dependence of$ R_{AA} $ is trivial in the current model of this study. We leave further development to future studies. Some qualitative results can be expected. Because the screening is a response of the medium to color charges inside a quarkonium, it takes time to form a screening cloud. For a fast-moving quarkonium, the screening cloud can never fully catch up with the quarkonium, and the screening is weakened [40]. Fast heavy qurakonia also have a higher chance to move to a region with lower temperature or even outside the fireball, which is known as the leakage effect [41]. Consequently, we expect a weaker suppression for fast-moving quarkonia.
Fast heating dissociation of ${ \Upsilon(1S) }$ in heavy ion collisions at RHIC
- Received Date: 2020-07-13
- Available Online: 2020-12-01
Abstract: By adopting the adiabatic assumption in the cooling process, we discuss a novel mechanism of