I. Why mesoscopic?
The mesoscopic description of physical systems bridges the gap between the
microscopic and macroscopic descriptions. The former (microscopic) approach
acounts for all constituents that make up the system and their fundamental
interactions. At the mesoscopic level, the individual constituents are
collectivelly described by means of a distribution function, defined over
phase space. The macroscopic description is oblivious of the phase-space
structure and instead relies on a further coarse-graining, replacing the
distribution function by a finite set of macroscopic observables, such as
energy density, fluid velocity, etc.
In principle, each level of description is accurate on its respective scale.
Highly non-equilibrium or nano-scale processes may represent the applicability
limit of the macroscopic description, while fundamentally quantum systems,
which are formed of a sufficiently small number of constituents, may prove
challenging for mesoscopic models.
In recent years, several arenas which bring together all three levels of
description have arised in a variety of systems. For example, multiscale
phenomena are ubiquitous challenges in astrophysics or cosmology,
such as the supernova explosions or large scale structure formation.
Below, the focus of the discussion will be on microscale systems, such
as the quark-gluon plasma formed in relativistic heavy ion colliders,
which also require a multiscale treatment.
II. Helical vortical effects
In the microscopic world, present day heavy ion colliders are able to create
the quark-gluon plasma (QGP), which exhibits properties which are remarkably
indicative of a nearly-perfect fluid behaviour. The macroscopic description of
the collective behaviour of the QGP relies on the correct realisation of
the transport laws that govern its evolution, which are derived from the
more fundamental mesoscopic description.
The vorticity of the QGP formed in non-central collisions can induce a net
polarisation, which was measured by the STAR collaboration at the Relativistic
Heavy Ion Collider (RHIC) at the Brookhaven National Laboratory (BNL).
This polarisation can be considered to be a signature of anomalous transport,
which traces its roots to the anomalous violation of classical conservation
laws via quantum fluctuations. When driven by local vorticity, these laws
are referred to as
Fig. 1: Kinematic tetrad corresponding to rigid rotation, comprised of the four-velocity
\(u = \Gamma(\partial_t + \Omega \partial_\varphi)\), acceleration
\(a = -\rho \Gamma^2 \partial_\rho\), vorticity
\(\omega = \Omega \Gamma^2 \partial_z\) and circular vector
\(\tau = -\Omega^3 \Gamma^5 (\rho^2 \Omega \partial_t + \partial_\varphi)\)
(Figure reproduced from Ref. [2]).
Table 1: Parity (\(+\) for even and \(-\) for odd) of the vector (\(V\)), axial (\(A\))
and helical (\(H\)) charge densities and (spatial) charge currents, as well as of the
kinematic vorticity \(\mathbf{\omega}\), under the C, P and T transformations (Table
reproduced from Ref. [1]).
QFT calculations reveal that the charge currents obey the following constitutive relations: \[ J_\ell^\alpha = Q_\ell u^\alpha + \sigma^\tau_\ell \tau^\alpha + \sigma^\omega_\ell \omega^\alpha. \] The first term, \(Q_\ell u^\alpha\), corresponds to the usual charge flow along the fluid macroscopic velocity. At large temperatures, the charge densities are given up to \(O(T^{-1})\) by: \[ Q_V = \frac{\mu_V T^2}{3} + \frac{4T \mu_A \mu_H}{\pi^2} \ln 2 + \frac{\mu_V(\mu_V^2 + 3\mu_A^2 + 3\mu_H^3)}{3\pi^2}, \] \[ Q_A = \frac{\mu_A T^2}{3} + \frac{4T \mu_H \mu_V}{\pi^2} \ln 2 + \frac{\mu_A(\mu_A^2 + 3\mu_H^2 + 3\mu_V^3)}{3\pi^2}, \] \[ Q_H = \frac{\mu_H T^2}{3} + \frac{4T \mu_V \mu_A}{\pi^2} \ln 2 + \frac{\mu_H(\mu_H^2 + 3\mu_V^2 + 3\mu_A^3)}{3\pi^2}, \] where the local temperature and chemical potentials are given in terms of their values on the rotation axis (denoted using the subscript \(0\)) and of the Lorentz factor \(\Gamma\): \[ T = \Gamma T_0, \qquad \mu_\ell = \Gamma \mu_{\ell,0}. \] The second term, \(\sigma^\tau_\ell \tau^\alpha\), represents a circular counterflow (provided \(\sigma^\tau_\ell Q_\ell > 0\)) along the circular vector \(\tau^\alpha\). Up to \(O(T^{-1})\), the circular conductivity \(\sigma^\tau_\ell\) is given by: \[ \sigma^\tau_V = \frac{\mu_V}{6\pi^2}, \qquad \sigma^\tau_A = \frac{\mu_A}{6\pi^2}, \qquad \sigma^\tau_H = \frac{\mu_H}{6\pi^2}. \] The term \(\sigma^\omega_\ell \omega^\alpha\) gives rise to current flow along the local vorticity \(\omega^\alpha\). While \(u^\alpha\) and \(\tau^\alpha\) vanish on the rotation axis, \(\omega^\alpha\) persists here, allowing the vortical effects to manifest themselves whenever a non-vanishing vorticity is present. The vortical conductivities are given up to \(O(T^{-1})\) terms by \[ \sigma^\omega_V = \frac{2\mu_H T}{\pi^2} \ln 2 + \frac{\mu_V \mu_A}{\pi^2}, \qquad \sigma^\omega_A = \frac{T^2}{6} + \frac{\mu_V^2 + \mu_A^2 + \mu_H^2}{\pi^2}, \qquad \sigma^\omega_H = \frac{2\mu_V T}{\pi^2} \ln 2 + \frac{\mu_H \mu_A}{\pi^2}. \] The robustness of the above transport laws at finite mass (and vanishing axial chemical potential) is discussed in Ref. [2], bearing in mind that when the mass is non-vanishing, the axial chemical potential can no longer be consistently introduced in the theory. In general, the massless expressions remain valid for \(M \lesssim T\). This can be seen in Fig. 2, where the charge density \(Q_\pm = Q_V \pm Q_H\) and the circular conductivity \(\sigma^\tau_\pm = \sigma^\tau_V \pm \sigma^\tau_H\) are represented with respect to the particle mass, \(M\).
Fig. 2: Ratio between the massive and massless values of the
charge density \(Q_\pm = Q_V \pm Q_H\) (left) and circular conductivity
\(\sigma^\tau_\pm = \sigma^\tau_V \pm \sigma^\tau_H\), represented with respect
to the mass \(M\), given in MeV (Figures reproduced from Ref. [2]).
Fig. 3: Dependence on (a) the distance \(\rho\) from the rotation axis, measured in fm,
and (b) on \(\Gamma\), of the energy densities obtained within the quantum
(\(E_\beta\), empty symbols and continuous lines) and relativistic kinetic theory
(\(E_F\), filled symbols and dashed lines) frameworks. The results are shown for various
values of the fermion particle mass \(M\). It can be seen that the quantum corrections
become dominant close to the SLS (Figures reproduced from Ref. [3]).
[1] V. E. Ambruș, M. N. Chernodub, Helical vortical effects, helical waves, and anomalies of Dirac fermions, arXiv:1912.11034 [hep-th].
[2] V. E. Ambruș, Helical massive fermions under rotation, JHEP08(2020)016.
[3] V. E. Ambruș, E. Winstanley, Exact solutions in quantum field theory under rotation, Book chapter in Strongly Interacting Matter Under Rotation. Edited by F. Becattini, J. Liao and M. Lisa Lecture Notes in Physics, vol. 987 (Springer, Cham, 2021).
III. Phase diagram of helically imbalanced QCD matter
Strongly interacting matter consists of quarks and gluons, which are
regarded in the theory of quantum chromodynamics (QCD) as elementary particles
interacting due to the colour charge. According to the standard
model of particles, a quark can have one of six flavours and are
arranged in pairs of charge \(+2/3\) and \(-1/3\), distributed
over three generations. The lightest generation consists of the
up (\(u\), charge \(+2/3\), mass \(\sim 2\) MeV) and down
(\(d\), charge \(-1/3\), mass \(\sim 5\) MeV) quarks.
The second generation, consists of the charm (\(c, 2/3, 1275\) MeV) and
strange (\(s, -1/3, 95\) MeV) quarks. The third generation consists of
the even heavier top (\(t, 2/3, 173\) GeV) and bottom (\(b, -1/3, 4.2\) GeV)
quarks. The heavy quarks from the second and third generations are unstable
and therefore much rarer (exotic) in normal conditions than the \(u\) and
\(d\) quarks.
In usual conditions, the quarks are bound together in colourless states,
called hadrons. Hadrons are typically mesons and baryons. The former
consist of a quark-anti-quark pair, such as the \(\pi^0 (u {\bar u} + d {\bar d})\),
\(\pi^+ (u {\bar d})\) and \(\pi^- (d {\bar u})\) mesons. The latter are formed
of three quarks (baryons) or three antiquarks (anti-baryons), the most common
baryons being the proton (\(uud\)) and neutron (\(udd\)). As either the temperature
or the density (modelled via the vector chemical potential, \(\mu_V\)) are increased,
the quarks are expected to break free from their hadronic prisons, giving rise
to the quark-gluon plasma.
An effective approach to model the QCD phase transition is to consider that
the quarks acquire a dynamical mass which is significantly heavier when the
quarks are confined into hadrons than when they are free. One such approaches
is the linear \(\sigma\) model with quarks, described using the following
Lagrangian:
\[
\mathcal{L}_{\rm LSM} = {\bar \psi} [i \gamma^\mu \partial_\mu -
g(\sigma +i\gamma^{5} \vec{\tau} \cdot \vec{\pi})]\psi +
\frac{1}{2} \left(\partial_\mu \sigma\partial^\mu \sigma
+ \partial_\mu \pi^0 \partial^{\mu}\pi^{0}\right) +
\partial_\mu \pi^+ \partial^{\mu}\pi^{-}
- V(\sigma,\vec{\pi}),
\]
where the potential \(V = \frac{\lambda}{4}(\sigma^2 + \vec{\pi}^2 - v^2)^2 - h\sigma\) achieves
its minimum when \(\langle \vec{\pi}\rangle = 0\) and \(\langle \sigma\rangle = f_\pi\).
The value of the free parameters appearing above are taken such that \(f_\pi = 93\) MeV is
equal to the pion decay constant. In the mean field approximation, the pion and sigma
fields are considered to be frozen at their classical expectation values, completely
neglecting their quantum fluctuations. In this case, the term \(g \langle \sigma\rangle = M\)
behaves like a dynamical quark mass, which is decided by the expectation value of the
pseudoscalar field \(\sigma\). Choosing \(g = 3.3\), the quark mass in the vacuum
is \(M = 307\) MeV, which is about one third of the mass of a
nucleon,
while the pion mass, given by \(m_\pi^2 = \lambda(\langle \sigma\rangle^2 - v^2)\),
takes the value \(m_\pi = 138\) MeV when \(v = 87.7\) MeV and \(\lambda = 19.7\). The
mass of the \(\sigma\) meson is \(m_\sigma = 600\) MeV.
Fig. 4: (top) Expectation value of \(\langle \sigma \rangle\), normalised with
respect to its vacuum value \(f_\pi\), at vanishing temperature, as a function of
the baryon chemical potential \(\mu_B = 3\mu_V\), for various values of the
helical chemical potential \(\mu_H\). (bottom) Phase diagram corresponding to the
chiral phase transition at \(T = 0\) (Figures reproduced from Ref. [4]).
Fig. 5: Phase diagram of the chiral phase transition at various values of
\(\mu_H\). The axes correspond to the temperature \(T\), normalised with respect
to the critical value \(T_{c,0} \simeq 146.5\) MeV, and vector chemical potential
\(\mu_V\). Each contour shown on the diagram corresponds to the phase transition
line for a fixed value of the helical chemical potential \(\mu_H\). The solid part of
these lines denote a first order phase transition, while the dotted parts correspond
to crossover transitions. The magenta self-duality line indicates the contour where
\(\mu_V = \mu_H\) (Figure reproduced from Ref. [4]).
[4] M. N. Chernodub, V. E. Ambruș, Phase diagram of helically imbalanced QCD matter, Phys. Rev. D 103 (2021) 094015.
IV. QFT on curved space
The Minkowski (flat-space) metric represents an idealisation in which the effect of
matter on the space-time fabric is completely neglected. In reality, any distribution
of mass-energy induces deviations from the flat-space limit. In particular, the
gravitational attraction is entirely due to the space-time curvature.
Taking into account the interaction between a quantum field and the gravitational
field can reveal interesting effects, of which probably the Hawking effect (predicting
the evaporation of black holes through thermal emission of radiation) remains the most
spectacular. Another important result refers to the so-called conformal anomaly
(also known as Weyl anomaly), due to which the trace of the stress-energy momentum
of a conformally-coupled massless fermion field aquires a non-vanishing value.
Of the curved backgrounds which have seen the most interest in the context of QCD is
the anti de Sitter (adS) space, due to the adS/CFT conjecture, which allows the
study of a conformally invariant, strongly interacting QFT via a dual weakly-coupled
gravitational system on the boundary of adS (one extra dimension compared to the
original system is required). To study the effect of rotation in such systems,
the results presented in this section focus on free fermions in rigid rotation
on adS. Ref. [6] discusses the behaviour of the fermion condensate at finite \(T\) and
\(\Omega\), but vanishing chemical potentials, as well as its surface and volume
integrals. Ref. [5] focusses more generally on anomalous transport.
The technique used for the analysis is point-splitting, starting from the vacuum
Feynman two-point function, \(S^F_{\rm vac}(x,x')\). When the rotation parameter is sufficiently
small and the SLS does not develop within the confinement of the adS boundary,
the thermal two-point function \(S^F_{\beta_0,\Omega}(x,x')\) can be obtained as [5,6]:
\[
S^F_{\beta_0,\Omega}(x,x') = \sum_{j = -\infty}^\infty (-1)^j e^{-j \beta_0 \Omega S^z}
S_{\rm vac}^F(\tau +ij \beta_0, \varphi + i j \beta_0 \Omega; x').
\]
The full details of the calculations can be found in Ref. [5]. It is remarkable that
in the large temperature limit and for massless fermions, the constitutive equations
obtained on the Minkowski space are exactly recovered. For example, the axial charge
current is \(J^\mu_A = \sigma^\omega_A \omega^\mu\), where \(\omega^\mu\) is the
local vorticity corresponding to rigid rotation, while the axial vortical conductivity
is
\[
\lim_{M\rightarrow 0} \sigma^\omega_{A} = -\frac{\omega^3 \cos^2\omega r}{2\pi^2 \Omega \Gamma^2}
\sum_{j = 1}^\infty \frac{(-1)^j \sinh\frac{\omega j \beta_0}{2} \sinh\frac{\Omega j \beta_0}{2}}
{(\sinh^2\frac{\omega j \beta_0}{2} - \omega^2 \overline{\rho}^2 \sinh^2\frac{\Omega j \beta_0}{2})^2}
= \frac{T^2}{6} + O(T^{-1}).
\]
It is remarkable that the fermion condensate (FC) aquires a non-vanishing value, which
persists even in the massless limit:
\[
\lim_{M \rightarrow 0} FC = -\frac{\omega^3}{2\pi^2} \cos^4\omega r \sum_{j = 1}^\infty
\frac{(-1)^j \cosh \frac{\omega j \beta_0}{2} \cosh \frac{\Omega j \beta_0}{2}}
{(\sinh^2 \frac{\omega j \beta_0}{2} + \cos^2 \omega r -
\omega^2 \overline{\rho}^2 \sinh^2\frac{\Omega j \beta_0}{2})^2} =
\frac{\omega^3}{4\pi^2} + O(T^{-1}).
\]
It can be seen that, at large temperature, the FC attains a constant value.
The volume and surface integrals of the FC, \(V_{\beta_0, \Omega}^{\rm FC}\) and
\(S_{\beta_0, \Omega}^{\rm FC},\) can also be computed [6] and are shown in
Fig. 6 below.
Fig. 6: Volume (a) and surface (b) integrals of the FC, represented
with respect to \((1 - \Omega^2 / \omega^2)^{-1}\), where
\(\Omega\) is the rotation parameter and \(\omega\) is the inverse adS radius
of curvature (the Ricci scalar is \(R = -12 \omega^2\)). Figure reproduced from Ref. [6].
[6] V. E. Ambruș, Fermion condensation under rotation on anti-de Sitter space, Acta Phys. Pol. B Proc. Suppl. 13 (2020) 199.
V. QFT between boundaries
Rigid rotation implies a linear increase of the azimuthal velocity with the
distance \(\rho\) to the rotation axis. At angular velocity \(\Omega\),
the speed of light surface (SLS), where the fluid velocity reaches the speed
of light, is equal to \(\Omega^{-1}\) (for the value typical for heavy
ion collisions, \(\Omega \simeq 10^{22}\,{\rm s}^{-1}\), the distance to the
SLS is \(c \Omega^{-1} = 0.3\ {\rm fm}\)). Close to the SLS, the macroscopic
quantities such as the energy density diverge. Surprisingly, quantum corrections
become dominant due to a higher order divergence close to the SLS. This can be
seen in Fig. 3 above. For these reasons, the study of rigidly rotating states
can only be self-consistently performed when the system is bounded such that
the SLS is excluded from the system. One example was the study of the
anti-de Sitter space discussed in Sec. IV above, which incorporates a natural
boundary. This section focuses on enclosing the usual Minkowski system inside
a cylindrical boundary, placed at distance \(R\) from the rotation axis.
While in hydrodynamics, the boundary conditions (b.c.s) are readily understood as
fixing the value (Dirichlet b.c.s) or the flux (Neumann b.c.s) of a given
quantity, in QFT, the b.c.s are formulated in a more abstract way, usually
at the level of the field operator itself. For the Klein-Gordon field,
both the Dirichlet- and Neumann-type boundary conditions are captured within
a more general scheme, known as Robin boundary conditions, which entail that
the field \(\Phi\) obeys on a boundary of normal vector \(n_\mu\) the following
equation:
\[ (A + B n^\mu \nabla_\mu) \Phi(x) = 0.\]
On the surface of the cylinder of radius \(R\), it is convenient to introduce
the parameter \(\Psi = A R / B\), such that the Robin boundary condition
becomes:
\[ R \frac{\partial \Phi}{\partial R} + \Psi \Phi = 0,
\]
where the Dirichlet boundary conditions are recovered when \(\Psi \rightarrow \infty\),
while the Neumann boundary conditions correspond to \(\Psi = 0\).
A challenging task is to establish the connection between the value of \(\Psi\)
and the properties of the quantum state in the vicinity of the boundary.
The quantum corrections (computed with respect to an equivalent classical,
relativistic kinetic theory system) also manifest themselves as a departure
from the perfect fluid form expected in global thermodynamic equilibrium.
Examples include the anomalous transport highlighted in Sec. II. The relativistic
equivalence between mass and energy make the concept of fluid four-velocity
itself ambiguous. For the specific case of the rigidly rotating system,
one natural choice for the four-velocity would be that corresponding to
rigid rotation, \(u = \Gamma(\partial_t + \Omega \partial_\varphi)\).
This choice establishes the so-called beta (or thermometer) frame.
However, this definition seems somewhat artificial or external,
since it relies on the a priori characterisation of the ensuing
quantum state. A more intrinsic definition of the four-velocity leads to
the so-called energy (or Landau) frame, when \(u_L^\mu\) is regarded as
an eigenvector of the stress-energy tensor, \(T^\mu{}_\nu u_L^\nu = E_L u_L^\mu\),
where the eigenvalue \(E_L\) is the Landau frame energy density. Solving this
equation, we find that the azimuthal velocity \(v_L = \rho u^{\varphi} / u^t\)
is related to the components of the stress-energy tensor via
\[
\frac{v_L}{1 + v_L^2} = \frac{\rho T^{t\varphi}}{T^{tt} + \rho^2 T^{\varphi\varphi}}.
\]
Evaluated on the boundary, the above relation can be taken as the starting
point for an interative procedure to find the value of \(\Psi\) that ensures
a given velocity \(v_b = v_L(\rho = R)\) on the boundary (more details can be found
in Ref. [7]). The left panel of Fig. 7 below shows how
the value of \(v_b\) depends on the Robin parameter \(\Psi\).
In the right panel of Fig. 7, the profiles of \(v_L\) are shown for selected
values of \(\Psi\), ranging from Neumann (\(\Psi = 0\)) to
Dirichlet (\(\Psi = \infty\)) boundary conditions.
Fig. 7: (a) Landau velocity on the boundary \(v_b\) for
various values of the inverse temperature \(\beta\) (expressed in
units of \(R / \hbar c\)) as a function of the Robin parameter
\(\Psi\). (b) Profiles of the Landau velocity for various values of
\(\Psi\), at inverse temperature \(\beta = 0.5 R\).
In both cases, the rotation paramter obeys \(R \Omega = 0.5 c\).
Figures reproduced from Ref. [7].