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

\jyear

2024 \Received2024/08/21\Accepted2024/11/18

\KeyWords

protoplanetary disks, planet–disk interactions, hydrodynamics

Modeling protoplanetary disk heating by planet-induced spiral shocks

Tomohiro Ono    11affiliation: School of Natural Science, Institute for Advanced Study, 1 Einstein Drive Princeton, New Jersey, 08540, USA \orcid0000-0001-8524-6939 Tatsuki Okamura    22affiliation: Department of Earth and Planetary Sciences, Institute of Science Tokyo, Ookayama, Meguro-ku, Tokyo, 152-8551, Japan Satoshi Okuzumi    22affiliationmark: \orcid0000-0002-1886-0880\altemailmark and Takayuki Muto33affiliation: Division of Liberal Arts, Kogakuin University, 1-24-2 Nishi-Shinjuku, Shinjuku-ku, Tokyo 163-8677, Japan [email protected]
Abstract

We investigate the heating of protoplanetary disks caused by shocks associated with spiral density waves induced by an embedded planet. Using two-dimensional hydrodynamical simulations, we explore the dependence of shock heating rates on various disk and planetary parameters. Our results show that the shock heating rates are primarily influenced by the planet’s mass and the disk’s viscosity, while being insensitive to the thermal relaxation rate and the radial profiles of the disk’s surface density and sound speed. We provide universal empirical formulas for the shock heating rates produced by the primary and secondary spiral arms as a function of orbital radius, viscosity parameter α\alpha, and planet-to-star mass ratio qq. The obtained formulas are accurate within a factor of a few for a moderately viscous and adiabatic disk with a planet massive enough that its spiral arms are strongly nonlinear. Using these universal relations, we show that shock heating can overwhelm viscous heating when the disk viscosity is low (α103\alpha\lesssim 10^{-3}) and the planet is massive (q103q\gtrsim 10^{-3}). Our empirical relations for the shock heating rates are simple and can be easily implemented into radially one-dimensional models of gas and dust evolution in protoplanetary disks.

1 Introduction

Protoplanetary disks are the birthplaces of planets. Their physical properties, such as temperature and ionization structures, influence the characteristics of the planets that form within them. The disk’s temperature structure is of particular interest because it plays a crucial role in determining the chemical composition of forming planets (e.g., Bitsch et al., 2019). For instance, the location of the snow line significantly influences the water content of protoplanets (e.g., Sato et al., 2016), although other internal processes can also affect their water budgets (Lichtenberg et al., 2019; Johansen et al., 2021).

Most models of disk thermal evolution (e.g., Bitsch et al., 2015; Mori et al., 2021) assume that disks are primarily heated by stellar irradiation and/or accretion heating. However, this assumption may not hold for disks that already harbor a massive planet. A planet embedded in a gas disk excites spiral density waves (e.g., Goldreich & Tremaine, 1979), which eventually shock as they propagate away from the planet (Goodman & Rafikov, 2001; Rafikov, 2002). These shocks generate entropy, causing secular heating of the disk gas (Rafikov, 2016). If the planet is massive and the disk is optically thick, this shock heating can raise the disk temperature (Richert et al., 2015; Lyra et al., 2016; Rafikov, 2016; Ziampras et al., 2020), potentially altering the chemical composition of the gas and dust remaining in the disk. Recent discoveries of protoplanetary disk substructures (e.g., Andrews, 2020; Bae et al., 2023) and isotopic dichotomies in meteorites (e.g., Kruijer et al., 2020; Kleine et al., 2020) suggest that giant protoplanets may have formed while their parent disks, including the solar nebula, were still abundant in solids. This implies that planet-induced shock heating could have impacted the composition of solar system bodies and exoplanets that formed from solids in the vicinity of giant planets.

Recently, Ziampras et al. (2020) conducted a systematic study of planet-induced shock heating using two-dimensional (2D) hydrodynamic simulations of disks with an embedded giant planet. They demonstrated that shock heating by a Jupiter-mass planet can significantly alter the disk’s temperature profile and even shift the location of the snow line. However, the disk temperature obtained directly from such simulations depends not only on shock heating but also on the disk’s radiative cooling (i.e., opacity) and accretion heating. Both the disk’s opacity and accretion heating rate are largely uncertain and can change with dust evolution (e.g., Oka et al., 2011; Kondo et al., 2023; Delage et al., 2023; Fukuhara & Okuzumi, 2024). To model planet-induced shock heating as independently as possible from these uncertainties, it is necessary to extract the shock heating rate from these simulations and analyze its dependence on the planet’s mass and distance from the planet.

In this paper, we model the planet-induced shock heating rate using 2D hydrodynamic simulations of disks with an embedded planet. From these simulations, we extract the radial profiles of the shock heating rates by directly measuring the jumps in the disk’s specific entropy across planet-induced shock waves. We discover universal scaling relations for the radial heating rate profiles from the simulations and provide simple analytic expressions for these universal relations.

The structure of this paper is as follows. In Section 2, we describe the simulation setups and analysis methods. Section 3 presents our simulation results and analysis. Section 4 discusses an application and caveats of our study, and Section 5 presents a summary.

2 Method

2.1 Basic equations

We consider a 2D gas disk surrounding a central star of mass MM_{\ast} and a planet of mass MpM_{\mathrm{p}} in a circular orbit with an orbital radius of rpr_{\mathrm{p}}. We adopt a frame centered at the position of the central star and co-rotating with the planet. The angular velocity of the planet in the inertial frame is Ωp=GM/rp3\Omega_{\mathrm{p}}=\sqrt{GM_{\ast}/r_{\mathrm{p}}^{3}}, where GG is the gravitational constant. In the co-rotating frame, the planet is located at (r,ϕ)=(rp,0)(r,\phi)=(r_{\mathrm{p}},0), where rr and ϕ\phi represent the distance from the central star and the azimuth, respectively.

The equations of continuity and motion in the co-rotating frame are given by

dΣdt=Σ𝒗,\frac{\mathrm{d}\Sigma}{\mathrm{d}t}=-\Sigma\nabla\cdot{\bm{v}}, (1)
Σd𝒗dt=p+ΠΣΦ+Σ(Ωp2𝒓2Ωp𝒏z×𝒗),\Sigma\frac{\mathrm{d}{\bm{v}}}{\mathrm{d}t}=-\nabla p+{\nabla\cdot\Pi}-\Sigma\nabla{\Phi}+\Sigma(\Omega_{\mathrm{p}}^{2}{\bm{r}}-2\Omega_{\mathrm{p}}{\bm{n}}_{z}\times{\bm{v}}), (2)

where Σ\Sigma is the surface density, 𝒗=(vr,vϕ){\bm{v}}=(v_{r},v_{\phi}) is the gas velocity, d/dt=/t+𝒗\mathrm{d}/\mathrm{d}t=\partial/\partial t+{\bm{v}}\cdot{\nabla} is the Lagrangian time derivative, pp is the vertically integrated pressure, Φ\Phi is the gravitational potential, 𝒓{\bm{r}} is the position vector, and 𝒏z{\bm{n}}_{z} is the unit vector perpendicular to the disk plane. The components of the viscous stress tensor Π\Pi are given by

Πij=Σν(vixj+vjxi23δij𝒗),\Pi_{ij}=\Sigma\nu\left(\frac{\partial v_{i}}{\partial x_{j}}+\frac{\partial v_{j}}{\partial x_{i}}-\frac{2}{3}\delta_{ij}{\nabla\cdot\bm{v}}\right), (3)

where ν\nu is the kinematic viscosity. We employ the α\alpha-viscosity model of Shakura & Sunyaev (1973) and give ν\nu as ν=αcsH\nu=\alpha c_{\rm s}H, where α\alpha is the viscosity parameter, csc_{\rm s} is the isothermal sound speed, H=cs/ΩKH=c_{\rm s}/\Omega_{\rm K} is the disk scale-height, and ΩK(r)=GM/r3\Omega_{\mathrm{K}}(r)=\sqrt{GM_{\ast}/r^{3}} is the Kepler angular velocity. We assume 104α10310^{-4}\leq\alpha\leq 10^{-3} (see section 2.4 for the complete parameter sets). The vertically integrated pressure pp can be expressed as

p=cs2Σ.p=c_{\mathrm{s}}^{2}\Sigma. (4)

The equation of energy is given by

dΣedt=\displaystyle\frac{\mathrm{d}\Sigma e}{\mathrm{d}t}= γΣe𝒗+12νΣ(Πrr2+2Πrϕ2+Πϕϕ2)\displaystyle-\gamma\Sigma e{\nabla\cdot{\bm{v}}}+\frac{1}{2\nu\Sigma}\left(\Pi_{rr}^{2}+2\Pi_{r\phi}^{2}+\Pi_{\phi\phi}^{2}\right)
+2νΣ9(𝒗)2+Qrelax,\displaystyle+\frac{2\nu\Sigma}{9}({\nabla\cdot{\bm{v}}})^{2}+Q_{\mathrm{relax}}, (5)

where e=p/[Σ(γ1)]e=p/[\Sigma(\gamma-1)] is the specific internal energy (i.e., internal energy per unit mass), γ\gamma is the adiabatic index (fixed to be 1.4 in this study), and QrelaxQ_{\mathrm{relax}} is the thermal relaxation term. We model QrelaxQ_{\mathrm{relax}} using the β\beta-relaxation (also known as the β\beta-cooling) prescription (e.g., Gammie, 2001), which enforces ee to relax to a reference value on timescale trelaxt_{\rm relax} (for details, see section 2.2). On the right-hand side of Equation (5), the first term represents compressional heating and cooling, while the second and third terms account for viscous heating. Due to the disk’s differential rotation, the term Πrϕ2/(νΣ)\Pi_{r\phi}^{2}/(\nu\Sigma) can, in principle, generate heat even in the absence of a planet. Since our focus is on heating from planet-induced spiral shocks, we subtract the local Keplerian velocity from 𝒗{\bm{v}} in the viscous heating terms to exclude heating due to Keplerian shear.

The net gravitational acceleration 𝒈gravΦ{\bm{g}}_{\mathrm{grav}}\equiv-\nabla{\Phi} is given by

𝒈grav\displaystyle{\bm{g}}_{\mathrm{grav}} =𝒈+𝒈p+𝒈in\displaystyle={\bm{g}}_{\star}+{\bm{g}}_{\mathrm{p}}+{\bm{g}}_{\mathrm{in}}
=GMr3𝒓GMp(re2+ϵ2)3/2𝒓eGMprp3𝒓p,\displaystyle=-\frac{GM_{*}}{r^{3}}{\bm{r}}-\frac{GM_{\mathrm{p}}}{(r_{\rm e}^{2}+\epsilon^{2})^{3/2}}{\bm{r}}_{\rm e}-\frac{GM_{\mathrm{p}}}{r_{\mathrm{p}}^{3}}{\bm{r}}_{\mathrm{p}}, (6)
𝒓e\displaystyle{\bm{r}}_{\rm e} =𝒓𝒓p,re=|𝒓e|\displaystyle={\bm{r}}-{\bm{r}}_{\mathrm{p}},\quad r_{\rm e}=|{\bm{r}}_{\rm e}|

where 𝒈{\bm{g}}_{\star}, 𝒈p{\bm{g}}_{\mathrm{p}}, and 𝒈in{\bm{g}}_{\mathrm{in}} represent the gravitational acceleration by the star, planet, and the indirect term, respectively, and 𝒓p{\bm{r}}_{\rm p} is the position vector of the planet. To avoid infinitely large acceleration in the vicinity of the planet, we adopt a smoothing length of ϵ=0.6rH\epsilon=0.6r_{\mathrm{H}}, where rH=(Mp/3M)1/3rpr_{\mathrm{H}}=(M_{\mathrm{p}}/3M_{\ast})^{1/3}r_{\mathrm{p}} is the planet’s Hill radius. Our choice of the smoothing length corresponds to 0.8H\approx 0.8H for q=103q=10^{-3}. Müller et al. (2012) show that choosing ϵ=(0.60.7)H\epsilon=(0.6\textrm{--}0.7)H best reproduces the vertically averaged gravitational force from the planet in a three-dimensional (3D) disk in the case of low-mass planets. More recently, the vertical averaging of the 3D planet potential is discussed by Brown & Ogilvie (2024), who present the form of 2D gravitational potential of a planet that matches 3D hydrodynamic calculations in the context of Type I migration.

2.2 Numerics

We numerically solve equations (1), (2), and (5) using the Athena++ code (Stone et al., 2020). We employ the Harten–Lax–van Leer–Contact approximate Riemann solver, the second-order piecewise linear method for spatial reconstruction, and the second-order Runge–Kutta method for time integration. We also adopt the orbital advection algorithm of Masset (2000).

The computational domain covers [rin,rout]=[0.4rp,9.88rp][r_{\mathrm{in}},r_{\mathrm{out}}]=[0.4r_{\mathrm{p}},9.88r_{\mathrm{p}}] in the radial direction and the full 2π2\pi in the azimuthal direction. Our fiducial runs use 896 and 1584 cells in the radial and azimuthal directions, respectively. We adopt equal spacing in the azimuthal direction and logarithmic spacing in the radial direction, ensuring that each cell has an aspect ratio of approximately unity. We also perform a supplementary, low-resolution simulation using 448×792448\times 792 cells to assess whether the finite resolution of our simulations impacts the entropy jumps at shocks (see section 3.1).

The initial surface density and sound speed, Σi\Sigma_{\rm i} and cic_{\rm i}, are given by

Σi(r)=Σ0(rrp)ξ,\Sigma_{\mathrm{i}}(r)=\Sigma_{0}\left(\frac{r}{r_{\mathrm{p}}}\right)^{\xi}, (7)

and

ci(r)=c0(rrp)ξ/23/4,c_{\mathrm{i}}(r)=c_{0}\left(\frac{r}{r_{\mathrm{p}}}\right)^{-\xi/2-3/4}, (8)

where Σ0\Sigma_{0} and c0c_{0} are the initial surface density and sound speed at r=rpr=r_{\mathrm{p}}, respectively. For a given initial surface density slope ξ\xi, the slope of the initial sound speed profile ξ/23/4-\xi/2-3/4 is determined to ensure that the viscous disk is initially in a steady state with a radially uniform mass accretion rate of M˙=3πνΣ\dot{M}=3\pi\nu\Sigma (Lynden-Bell & Pringle, 1974). We adopt c0=0.05vK(rp)c_{\mathrm{0}}=0.05v_{\mathrm{K}}(r_{\mathrm{p}}), where vK(r)v_{\rm K}(r) is the local Keplerian velocity. Our grid resolves the local disk scale height with at least 12 cells for r0.55rpr\geq 0.55r_{\mathrm{p}}.

The initial gas velocity 𝒗i=(vri,vϕi){\bm{v}}_{\rm i}=(v_{r\mathrm{i}},v_{\phi\mathrm{i}}) is given by

vri=3αci22rΩK,v_{r\mathrm{i}}=-\frac{3\alpha c_{\mathrm{i}}^{2}}{2r\Omega_{\mathrm{K}}}, (9)
vϕi(r)=rΩi(r)=vK2(r)ci2(r),v_{\phi\mathrm{i}}(r)=r\Omega_{\mathrm{i}}(r)=\sqrt{v_{\mathrm{K}}^{2}(r)-c_{\mathrm{i}}^{2}(r)}, (10)

where Ωi\Omega_{\mathrm{i}} is the initial angular velocity of the disk. The initial radial velocity is consistent with steady accretion with vr=3ν/(2r)v_{r}=-{3\nu}/(2r) (Lynden-Bell & Pringle, 1974). The initial azimuthal velocity vϕiv_{\phi\mathrm{i}} ensures the balance among the central star’s gravity, the centrifugal force, and the pressure gradient force.

The relaxation term QrelaxQ_{\mathrm{relax}} is given by

Qrelax=Σ(eei)Ωiβ,Q_{\mathrm{relax}}=-\frac{\Sigma(e-e_{\mathrm{i}})\Omega_{\mathrm{i}}}{\beta}, (11)

where β\beta is a parameter related to the thermal relaxation timescale trelaxt_{\rm relax} by trelax=βΩi1t_{\rm relax}=\beta\Omega_{\mathrm{i}}^{-1} and ei(r)=ci2(r)/(γ1)e_{\mathrm{i}}(r)=c_{\mathrm{i}}^{2}(r)/(\gamma-1) is the initial internal energy per unit mass. The case with β=0\beta=0 represents locally isothermal disks, while the case with β=\beta=\infty represents adiabatic disks. In this study, we take β\beta to be constant throughout the disk.

The inner and outer boundaries of the computational domain are fixed to their initial values, while a periodic boundary condition is imposed on the azimuthal boundaries. To reduce wave reflection at the radial boundaries, we introduce wave-damping zones at rin<r<1.25rinr_{\mathrm{in}}<r<1.25r_{\mathrm{in}} and 0.8rout<r<rout0.8r_{\mathrm{out}}<r<r_{\mathrm{out}}, following Val-Borro et al. (2006). In these zones, the gas velocity is enforced to return to the initial velocity 𝒗i{\bm{v}}_{i} on a timescale of 1/η1/\eta, according to

𝒗t=η(𝒗𝒗i),\frac{\partial{\bm{v}}}{\partial t}=-\eta({\bm{v}}-{\bm{v}}_{\mathrm{i}}), (12)

where

η=min(x,1)ΩK\eta=\mathrm{min}(x,1)\Omega_{\mathrm{K}}\\ (13)

with

x={(1.25rinr)/rp,rin<r<1.25rin,2.5(r0.8rout)/rp,0.8rout<r<rout.x=\begin{cases}(1.25r_{\mathrm{in}}-r)/r_{\mathrm{p}},&r_{\mathrm{in}}<r<1.25r_{\mathrm{in}},\\ 2.5(r-0.8r_{\mathrm{out}})/r_{\mathrm{p}},&0.8r_{\mathrm{out}}<r<r_{\mathrm{out}}.\end{cases}\\ (14)

At the beginning of each simulation, the planet mass is gradually increased from zero to the final value MpM_{\rm p} according to

M~p(t)={Mpsin2[12π(tτins)],t<τins,Mp,tτins,\tilde{M}_{\mathrm{p}}(t)=\begin{cases}M_{\mathrm{p}}\sin^{2}{\left[\frac{1}{2}\pi\left(\frac{t}{\tau_{\mathrm{ins}}}\right)\right]},&t<\tau_{\mathrm{ins}},\\ M_{\mathrm{p}},&t\geq\tau_{\mathrm{ins}},\end{cases}\\ (15)

where M~p(t)\tilde{M}_{\mathrm{p}}(t) is the planetary mass at time tt. We set τins=50×2π/ΩK(rp)\tau_{\mathrm{ins}}=50\times 2\pi/\Omega_{\mathrm{K}}(r_{\mathrm{p}}).

Our simulations adopt code units such that rp=Ωp=Σ0=1r_{\rm p}=\Omega_{\rm p}=\Sigma_{\rm 0}=1, where ΩpΩK(rp)\Omega_{\rm p}\equiv\Omega_{\mathrm{K}}(r_{\mathrm{p}}) is the orbital frequency of the planet. Since we take QrelaxQ_{\rm relax} to be proportional to Σ\Sigma, equations (1), (2), and (5) are invariant under scaling ΣCΣ\Sigma\to C\Sigma, where CC is an arbitrary constant. Therefore, the scaling factor Σ0\Sigma_{0} in equation (7) can be chosen arbitrarily in our simulations.

2.3 Analysis of shock wave heating

A planet embedded in a disk excites spiral density waves that travel through the disk away from the planet and eventually shock (Goodman & Rafikov, 2001; Rafikov, 2002). Because these waves rotate at the planet’s orbital frequency, while the disk gas rotates differentially, they sweep through the disk gas at a given orbital radius rr with a frequency of |ΩK(r)Ωp||\Omega_{\rm K}(r)-\Omega_{\rm p}|. In this study, we quantify the shock wave heating rate by measuring the jump in the disk gas entropy across the planet-induced spiral waves.

The entropy change is a key physical quantity that characterizes the magnitude of gas heating at a shock. We denote the pre-shock gas density by Σ1\Sigma_{1}, pressure by p1p_{1}, and specific entropy by s1=p1/Σ1γs_{1}=p_{1}/\Sigma_{1}^{\gamma}, and the post-shock values by Σ2(>Σ1)\Sigma_{2}(>\Sigma_{1}), p2(>p1)p_{2}(>p_{1}), and s2=p2/Σ2γ(>s1)s_{2}=p_{2}/\Sigma_{2}^{\gamma}(>s_{1}), respectively. We define the dimensionless specific entropy jump δ\delta as

δs2s11=p2/Σ2γp1/Σ1γ1.\delta\equiv\frac{s_{2}}{s_{1}}-1=\frac{p_{2}/\Sigma_{2}^{\gamma}}{p_{1}/\Sigma_{1}^{\gamma}}-1. (16)
Refer to caption
Figure 1: Left panel: Snapshot of the 2D disk surface density distribution in the quasi-steady state from RUN 1. The white dashed and dotted lines mark the primary and secondary spiral arms, respectively. The black line indicates a streamline passing through the point (r,ϕ)=(2rp,0)(r,\phi)=(2r_{\mathrm{p}},0). Right panel: Zoom-in of the vicinity of the streamline shown in the left panel. In the planet’s rest frame, the gas flow on the streamline is in the direction of decreasing ϕ\phi.

The specific entropy jump is directly measurable from numerical simulations by tracing the motion of gas fluid elements. We take a snapshot of a simulation after the system reaches a quasi-steady state, which occurs after 1000 orbits of the planet for models with α=103\alpha=10^{-3} and 3000 orbits for those with α=103.5\alpha=10^{-3.5} or 10410^{-4}. On the snapshot, we draw gas streamlines originating from different orbital distances (see the right panel of figure 2 in section 3 for an example of such streamlines) and measure the entropy along each of them. The entropy along the streamlines is calculated using bilinear interpolation of the 2D grid data. Shocks manifest as jumps in the entropy profile. We calculate the δ\delta value of each shock from the minimum and maximum entropy values immediately before and after the shock. By compiling the specific entropy jumps from different streamlines, we obtain δ\delta as a function of rr. We discard small entropy jumps where δ\delta is less than 20% of the largest shock’s value.

Our analysis excludes the region in the vicinity of the planet, specifically 0.8rp<r<1.2rp0.8r_{\mathrm{p}}<r<1.2r_{\mathrm{p}}, where a deep planet-induced gap prevents reliable measurements of entropy changes associated with shocks. In practice, shock-induced heating is not significant in a deep planet-induced gap where the optical depth is small and radiative cooling is rapid (see Ziampras et al., 2020).

The specific entropy jump can be translated into the heating rate QshQ_{\mathrm{sh}} of the gas per unit area by (see Rafikov, 2016)

Qsh=p1Δt(γ1)δ,Q_{\mathrm{sh}}=\frac{p_{1}}{\Delta t(\gamma-1)}\delta, (17)

where Δt=2π/|ΩK(r)Ωp|\Delta t=2\pi/|\Omega_{\mathrm{K}}(r)-\Omega_{\mathrm{p}}| is the time interval between two consecutive shock crossings for a gas particle orbiting at radius rr. As we show in the following section, the planet in our simulations induces multiple spirals both inside and outside of its orbit, with each spiral forming a shock. When multiple spiral arms exist along a single streamline, the net heating rate is proportional to the sum of the δ\delta values at all these arms. Using equation (4), the shock heating rate can also be written as

Qsh=Σcs2Δt(γ1)δ=Σe2π|ΩKΩp|δ,Q_{\rm sh}=\frac{\Sigma c_{\rm s}^{2}}{\Delta t(\gamma-1)}\delta=\frac{\Sigma e}{2\pi}|\Omega_{\rm K}-\Omega_{\rm p}|\delta, (18)

where we have omitted the subscript 1 for the preshock quantities. When there are multiple planet-induced shocks, δ\delta in equation (18) should be regarded as the sum of the specific entropy jumps at the individual shocks. In section 3.1, we numerically confirm that equation (18) accurately describes the azimuthally averaged rate of heat generation by the spiral shocks.

2.4 Parameter sets

We present 11 runs with different values of (q,ξ,α,β)(q,\xi,\alpha,\beta) as listed in table 1, where q=Mp/Mq={M}_{\rm p}/M_{\ast} is the planet-to-star mass ratio. We refer to RUN 1 as the fiducial model. For all runs except RUN 6, the planet mass mass is larger than the thermal mass qth=(c0/vK(rp))3104q_{\rm th}=(c_{0}/v_{\rm K}(r_{\rm p}))^{3}\approx 10^{-4}, which is the minimum planet mass required for planet-induced waves to be strongly nonlinear and form a shock immediately after launching (Lin & Papaloizou, 1993; Rafikov, 2002). Our simulations represent moderately adiabatic cases with 10β102.510\leq\beta\leq 10^{2.5}. For β=1\beta=1, we cannot accurately measure the δ\delta values at the shocks because the strong cooling largely erases the entropy jumps. However, this is not a significant issue, as shock heating would have little effect on the disk temperature in such rapidly cooling disks (see also Ziampras et al., 2020). In more adiabatic cases with β>102\beta>10^{2}, the disk’s specific internal energy (or temperature) can significantly exceed the initial value, leading to numerical instability at the inner boundary, where the specific internal energy is fixed to the initial value. Our simulation is unstable with α=103\alpha=10^{-3} and β102.5\beta\geq 10^{2.5}, but is barely stable with α=103.5\alpha=10^{-3.5} and β=102.5\beta=10^{2.5} (RUN 11).

Table 1: List of parameter sets
RUN qq ξ\xi α\alpha β\beta
RUN 1 10310^{-3} 1.0-1.0 10310^{-3} 10210^{2}
RUN 2 10310^{-3} 0.5-0.5 10310^{-3} 10210^{2}
RUN 3 10310^{-3} 1.5-1.5 10310^{-3} 10210^{2}
RUN 4 103.510^{-3.5} 1.0-1.0 10310^{-3} 10210^{2}
RUN 5 10410^{-4} 1.0-1.0 10310^{-3} 10210^{2}
RUN 6 104.510^{-4.5} 1.0-1.0 10310^{-3} 10210^{2}
RUN 7 10310^{-3} 1.0-1.0 103.510^{-3.5} 10210^{2}
RUN 8 10310^{-3} 1.0-1.0 10410^{-4} 10210^{2}
RUN 9 10310^{-3} 1.0-1.0 10310^{-3} 1010
RUN 10 10310^{-3} 1.0-1.0 10310^{-3} 101.510^{1.5}
RUN 11 10310^{-3} 1.0-1.0 103.510^{-3.5} 102.510^{2.5}

3 Results

3.1 The fiducial case

Refer to caption
Figure 2: Distribution of the surface density (dashed line) and specific entropy s=pΣγs=p\Sigma^{-\gamma} (red line) of the gas along the streamline shown in figure 1. The jumps in surface density and entropy observed at ϕ0.25π\phi\sim 0.25\pi and 0.75π\sim-0.75\pi correspond to the shocks associated with the primary and secondary arms, respectively (see also the right panel of figure 1).

We begin by presenting the results from the fiducial run (RUN 1) to illustrate our shock wave analysis. The left panel of figure 1 shows a snapshot of the 2D gas surface density profile in the quasi-steady state (t=1000t=1000 planetary orbits) from RUN 1. The planet creates a deep disk gap in its vicinity (r0.8r\sim 0.81.2rp1.2r_{\rm p}). The planet also excites spiral waves both interior to and exterior to its orbit. As discussed in previous studies (Zhu et al., 2015; Fung & Dong, 2015; Bae et al., 2017; Bae & Zhu, 2018), a planet excites more than one spiral arm in both the inner and outer disks. Following the previous studies, we refer to the arms directly attached to the planet as the primary arms and to the strongest arms that do not connect to the planet as the secondary arms. In the left panel of figure 1, the primary and secondary arms are marked by the dashed and dotted lines, respectively.

As described in section 2.3, we draw streamlines at different orbits and measure the entropy jumps across the arms. The black line in the left panel of figure 1 shows the streamline passing through the point (r,ϕ)=(2rp,0)(r,\phi)=(2r_{\mathrm{p}},0) in this snapshot (see also the zoom-in of the vicinity of the streamline in the right panel of figure 1). Because the system has already reached a quasi-steady state, the streamlines on this snapshot approximately follow the trajectories of gas fluid elements. Figure 2 shows how the gas surface density and entropy of a fluid element along this streamline vary in the direction of the flow in the planet’s rest frame (negative ϕ\phi). The surface density of the element abruptly increases at the shock due to shock compression and then returns to the background value due to expansion. The entropy also increases at the shocks due to shock dissipation and then decreases due to cooling relaxation (the QrelaxQ_{\rm relax} term in equation (5)). After shock crossing, the change in entropy of the moving parcel decays on the timescale of tc\sim t_{c}, which is approximately 100 times the local Keplerian time. A close inspection of figure 2 shows that the measured entropy profile exhibits a small bump at ϕ0.1π\phi\sim 0.1\pi. This is a numerical artifact that arises when the streamline crosses a radial cell boundary and enters an adjacent outer cell with slightly higher entropy. However, this artificial bump does not affect our measurement of the entropy jump at the primary arm’s shock, located at ϕ/π0.2\phi/\pi\sim 0.2.

Refer to caption
Figure 3: Radial profiles of the specific entropy jump δ\delta (equation (16)) for the primary and secondary arms (dashed and dotted lines, respectively) from RUN 1. The solid line shows the sum of the two jumps, representing the net shock heating rate. Our shock analysis excludes the shaded region of 0.8<r/rp<1.20.8<r/r_{\mathrm{p}}<1.2, where the deep planet-carved gap prevents an accurate measurement of δ\delta using streamlines.

The streamline analysis illustrated above provides the dimensionless specific entropy jumps δ\delta (equation (16)) at the primary and secondary shocks at different rr, except in the gapped region 0.8<r/rp<1.20.8<r/r_{\rm p}<1.2, which is excluded from the analysis. These values are plotted in figure 3 as a function of rr. The primary inner and outer arms yield a non-zero δ\delta at r/rp=0.8r/r_{\rm p}=0.8 and 1.2, respectively, implying that both have already developed into shocks before propagating out of the gapped region. In contrast, the secondary inner and outer arms exhibit detectable entropy jumps at r/rp0.7r/r_{\rm p}\lesssim 0.7 and 2.0\gtrsim 2.0, respectively, indicating that the secondary arms form shocks only after traveling considerable distances away from the planet. Interestingly, at large |rrp||r-r_{\rm p}|, the δ\delta values of the secondary shocks surpass those of the primary shocks, meaning that disk heating by the secondary arms dominates over that by the primary ones. This is because the secondary arms have not experienced dissipation until they form shocks. The potential significance of the secondary shocks in disk heating is further discussed in section 4.1.

In the log–log plot of δ\delta versus rr (figure 3), the radial profiles of δ\delta for the inner and outer primary shocks appear as straight lines. This indicates that both approximately follow power laws. The same holds true for the fully developed secondary outer shock at r/rp>2r/r_{\rm p}>2. We cannot confirm such a trend for the secondary inner shock because the inner computational domain is too narrow for the secondary shock to fully develop. We provide empirical formulas for the entropy jumps in section 3.3.

Refer to caption
Figure 4: Radial profile of the azimuthally averaged specific internal energy ee, normalized by e0ei(rp)e_{0}\equiv e_{\rm i}(r_{\rm p}), from RUN 1 in the quasi-steady state (solid line), compared with the prediction that assumes thermal equilibrium Qsh+Qrelax=0Q_{\rm sh}+Q_{\rm relax}=0 (equation (19); dashed lines). The dotted line shows the initial profile e=eie=e_{\rm i}. See the caption of figure 3 for the shaded region. Note that viscous heating due to Keplerian shear is excluded from the simulation (see section 2.1).

Before proceeding to a parameter study, we verify that the entropy jumps produced by the spiral shocks indeed cause disk heating, as predicted by equation (18). Figure 4 shows the radial profile of the azimuthally averaged specific internal energy from RUN 1 in the quasi-steady state. We expect that this internal energy profile is determined by the balance between shock heating and β\beta thermal relaxation (note that our simulations do not include viscous heating from Keplerian shear). Solving the thermal equilibrium equation Qsh+Qrelax=0Q_{\rm sh}+Q_{\rm relax}=0 for ee, the internal energy profile in the steady state is predicted as

eequil=ei(1β|1Ωp/Ω(r)|2πδ)1.e_{\rm equil}=e_{\rm i}\left(1-\frac{\beta|1-\Omega_{\rm p}/\Omega(r)|}{2\pi}\delta\right)^{-1}. (19)

We calculate eequile_{\rm equil} using the radial δ\delta profile obtained from the streamline analysis presented above. The result is overplotted in figure 4. We find that eequile_{\rm equil} reproduces the azimuthally averaged ee in the steady state to within an accuracy of 5\approx 5%.

Refer to caption
Figure 5: Radial profiles of the specific entropy jumps for the primary and secondary arms (dashed and dotted lines, respectively) from a low-resolution version of RUN 1 (gray lines), compared with those from RUN 1 at standard resolution shown in figure 3 (black lines).

We also briefly examine the impact of numerical resolution on our measurements of entropy jumps. Here, we use a low-resolution version of RUN 1, with 448×792448\times 792 cells, resulting in half the resolution in both the radial and azimuthal directions. Figure 5 compares the radial profiles of δ\delta for the primary and secondary arms from the low-resolution simulation with those from RUN 1 at standard resolution (896×1584896\times 1584 cells). We find that the low-resolution run reproduces the δ\delta values for relatively large jumps (δ0.005\delta\gtrsim 0.005) within 20% accuracy. The low-resolution run underestimates the smaller entropy jump at the secondary outer arm, but the error remains within \approx 30%. Based on these comparisons, we expect that the simulations presented in this study produce δ\delta values with a resolution-related error of less than 20–30%.

3.2 Parameter survey

Refer to caption
Figure 6: Radial profiles of the azimuthally averaged surface density in the quasi-steady state from RUNs 1–10. The upper left, upper right, lower left, and lower right panels compare the results from runs with different planetary masses qq (RUNs 1 and 4–6), initial surface density slopes ξ\xi (RUNs 1–3), viscosity parameters α\alpha (RUNs 1, 7, and 8), and thermal relaxation timescales β\beta (RUNs 1, 9, 10), respectively.

We now explore how our simulation results depend on the model parameters. Figure 6 shows the radial profiles of the azimuthally averaged surface density in the quasi-steady state from RUNs 1–10. It can be seen that different values of the planet mass qq, surface density slope ξ\xi, viscosity parameter α\alpha, and thermal relaxation rate β\beta result in varying surface density profiles. However, as we will see below, the radial profiles of the entropy jumps at the spiral shocks are surprisingly similar among the models when scaled properly.

The upper left and right panels of figure 7 show the radial profiles of δ\delta for the primary and secondary arms, respectively, from runs with different planet masses (RUNs 1 and 4–6). Their behavior is qualitatively similar to that observed in RUN 1 (see the previous subsection). Specifically, regardless of the planet’s mass, the δ\delta values of the primary arms outside the gap region decreases monotonically with increasing |rrp||r-r_{\rm p}|, while the δ\delta values of the secondary arms peak after traveling some distance from the planet. Once the secondary shock fully develops, its δ\delta value exceeds that of the primary shock at the same location. The only significant difference is that the magnitude of δ\delta scales approximately linearly with qq. To see this, in the lower panels of figure 7, we plot the entropy jumps normalized by the planetary mass, δ/q\delta/q, as a function of rr from the four runs. We observe that the radial profiles of the normalized entropy jump almost coincide, particularly in the cases of q=104q=10^{-4}10310^{-3}. The trend shifts slightly in the lowest-mass case of q=104.5q=10^{-4.5}, potentially because the shocks for this case are too weak to resolve accurately.

Figure 8 shows the radial profiles of δ\delta for the primary and secondary arms from runs with different values of ξ\xi (RUNs 1–3). The δ\delta profile for the primary arms is nearly independent of ξ\xi, despite the background surface density profile depending on ξ\xi. The same applies to the δ\delta profile for the secondary arms. The only notable trend is that the location where the secondary arms form a shock (i.e., where δ\delta becomes detectable) depends slightly on ξ\xi. Specifically, as ξ\xi decreases, the shock formation position for the inner secondary arm shifts closer to the planet, while for the outer arm, it shifts farther. This dependence on ξ\xi likely reflects the initial radial sound velocity profile, cirξ3/4c_{i}\propto r^{-\xi-3/4} (see equation (8)). As ξ\xi decreases, the sound speed at r<rpr<r_{\rm p} (r>rpr>r_{\rm p}) decreases (increases), causing the secondary arm to form a shock after traveling a shorter (longer) distance from the planet.

The upper panels of figure 9 show the same as figure 8 but from runs with different values of α\alpha (RUNs 1, 7, and 8). In most cases, δ\delta crudely scales as α\sqrt{\alpha}, as demonstrated in the lower panels of figure 9. This indicates that δ\delta from our simulations depends weakly on the assumed viscosity. A notable exception is the secondary shock for α=104\alpha=10^{-4}, whose δ\delta profile approximately follows that of the secondary shock for α=103.5\alpha=10^{-3.5}.

Figure 10 shows that varying β\beta over 10β10210\leq\beta\leq 10^{2} has little effect on δ\delta. We note only that the peak of the δ\delta distribution for the secondary outer arm shifts slightly outward as we vary β\beta from 101.510^{1.5} to 10210^{2}. This shift occurs because a higher β\beta results in an overall hotter disk with a higher sound speed, causing the secondary arm to form a shock farther from the planet.

Refer to caption
Figure 7: Upper panels: radial profiles of the specific entropy jumps δ\delta at the primary and secondary arms (left and right panels, respectively) from runs with different planetary masses qq (RUNs 1 and 4–6). Lower panels: same as the upper panels, but show the entropy jumps normalized by the planetary mass, δ/q\delta/q.
Refer to caption
Figure 8: Same as the upper panels of figure 7, but from runs with different initial surface density slopes ξ\xi (RUNs 1–3).
Refer to caption
Figure 9: Upper panels: Same as the upper panels of figure 7, but from runs with different viscosity parameters α\alpha (RUNs 1, 7, and 8). Lower panels, same as the upper panels, but show the entropy jumps normalized by α\sqrt{\alpha}.
Refer to caption
Figure 10: Same as the upper panels of figure 7, but from runs with different thermal relaxation timescales β\beta (RUNs 1, 9, and 10).

3.3 Empirical formulas for the entropy jumps

In the previous subsection, we have observed that δ\delta for the primary and secondary arms follows universal relations that only depend on qq, α\alpha, and r/rpr/r_{\rm p}. Here, we provide empirical formulas for the universal relations.

Refer to caption
Figure 11: Universal relations for the specific entropy jumps for the primary arms. The dashed curves show the radial profiles of δ/qα\delta/q\sqrt{\alpha} for the primary inner and outer arms from all simulation runs. The black solid lines show the empirical formulas proposed in this study (equation (20)). The yellow shading surrounding the solid lines indicates a factor of 2 deviation from the formulas. Our shock analysis excludes the gray-shaded region of 0.8<r/rp<1.20.8<r/r_{\mathrm{p}}<1.2.

For the primary arms, we assume that δ\delta takes the power-law form

δ=10f0qα|r/rp|f1,\delta=10^{f_{0}}q\sqrt{\alpha}\cdot|r/r_{\mathrm{p}}|^{f_{1}}, (20)

where f0f_{0} and f1f_{1} are fitting parameters. We searched for the set of (f0,f1)(f_{0},f_{1}) that best reproduces the δ\delta profiles for each of the primary inner and outer arms, using the data from all runs except for the lowest planet mass case (RUN 6; q=104.5q=10^{-4.5}) and the low viscosity case (RUN 8; α=104\alpha=10^{-4}), which appear to be slightly out of trend. The obtained best-fit parameters for the primary inner and outer arms are (f0,f1)=(4.22,10.02)(f_{0},f_{1})=(4.22,10.02) and (f0,f1)=(3.87,7.37)(f_{0},f_{1})=(3.87,-7.37), respectively. Figure 11 demonstrates that the best-fit formulas reproduce the results form all runs except RUNs 6 and 8 to within a factor of 2. Assuming radiative equilibrium where QshσT4Q_{\rm sh}\propto\sigma T^{4}, with σ\sigma being the Stefan–Boltzmann constant, a factor of 2 error in δ\delta translates into an approximately 20% error in the disk interior temperature TT. For RUNs 6 and 8, our empirical formulas are less accurate, with a relative error of up to a factor of 3. The smaller-than-expected entropy jumps observed in RUN 6 are potentially due to the inefficient nonlinearity of the planet-induced waves, i.e., q<qthq<q_{\rm th}, in this run (see section 2.4). We note that the resolution-related errors in the measured δ\delta values are less than 20–30% (see section 3.1) and are therefore smaller than the errors arising from the fitting.

We also provide an empirical formula for δ\delta for the secondary outer arm, which has been found to fully develop into a shock within our computational domain. We assume that the secondary arm’s δ\delta also follows the form given by (20) but with a cutoff at small |rrp||r-r_{\rm p}|. By fitting our simulation results, we find that δ\delta for the secondary outer arm scales as r4.79r^{-4.79} at large radial distances. We propose the formula

δ=103.7qα(rrp)4.79[(r/rp10.95)12+1]1,\delta=10^{3.7}q\sqrt{\alpha}\biggl{(}\dfrac{r}{r_{\mathrm{p}}}\biggr{)}^{-4.79}\left[\biggl{(}\dfrac{r/r_{\mathrm{p}}-1}{0.95}\biggr{)}^{-12}+1\right]^{-1}, (21)

where the last factor represents the cutoff. This formula reproduces our simulation results to within a factor of a few (figure 12).

Refer to caption
Figure 12: Same as figure 11, but for the secondary outer arm.

4 Discussion

4.1 Comparison between the shock and viscous heating rates

As noted in section 2.1, our simulations do not include viscous heating from the disk’s Keplerian shear. In reality, however, both Keplerian shear and planet-induced shocks contribute to disk heating. Here, we use the analytic expressions for the specific entropy jump δ\delta obtained in this study to predict when and where shock heating may dominate over viscous heating.

The viscous heating rate is given by (Lynden-Bell & Pringle, 1974; Pringle, 1981)

Qvis=94νΣΩK2=9(γ1)4αΣeΩK.Q_{\rm vis}=\frac{9}{4}\nu\Sigma\Omega_{\rm K}^{2}=\frac{9(\gamma-1)}{4}\alpha\Sigma e\Omega_{\rm K}. (22)

The ratio between the shock heating rate QshQ_{\rm sh} (equation (18)) and QvisQ_{\rm vis} can be written as

QshQvis\displaystyle\frac{Q_{\rm sh}}{Q_{\rm vis}} =\displaystyle= 29π(γ1)|1ΩpΩK(r)|δα\displaystyle\frac{2}{9\pi(\gamma-1)}\left|1-\frac{\Omega_{\rm p}}{\Omega_{\rm K}(r)}\right|\frac{\delta}{\alpha} (23)
\displaystyle\approx 0.18|1(rrp)3/2|δα.\displaystyle 0.18\left|1-\left(\frac{r}{r_{\rm p}}\right)^{3/2}\right|\frac{\delta}{\alpha}.

Notably, the ratio depends only on r/rpr/r_{\rm p} and δ/α\delta/\alpha. Furthermore, except at rrpr\gg r_{\rm p}, the ratio is determined solely by δ/α\delta/\alpha.

Refer to caption
Figure 13: Ratio between the shock and viscous heating rates, Qsh/QvisQ_{\rm sh}/Q_{\rm vis} (Equation (23)), as a function of r/rpr/r_{\rm p} for different values of the dimensionless planetary mass qq and viscosity α\alpha. Here, the shock heating rate accounts for the primary and secondary shocks at r>rpr>r_{\rm p} and only the primary shock at r<rpr<r_{\rm p}, using the empirical formulas for δ\delta given by equations (20) and (21). Note that the formulas are validated only outside the gray-shaded region of 0.8<r/rp<1.20.8<r/r_{\mathrm{p}}<1.2, which is excluded from our shock analysis.

Our empirically obtained relation δqα\delta\propto q\sqrt{\alpha} suggests that the heating rate ratio Qsh/QvisQ_{\mathrm{sh}}/Q_{\mathrm{vis}} scales as q/αq/\sqrt{\alpha}. Figure 13 illustrates the radial distribution of Qsh/QvisQ_{\mathrm{sh}}/Q_{\mathrm{vis}} for q=104q=10^{-4} and 10310^{-3}. Here, the shock heating rate accounts for the contributions from the primary inner and outer arms, as well as the secondary outer arm (equations (11) and (12), respectively). We find that shock heating by a planet with mass q=103q=10^{-3} can overwhelm viscous heating if α<103\alpha<10^{-3}. Our result aligns with the findings by Ziampras et al. (2020), who showed that the shocks induced by a Jupiter-mass planet orbiting a solar-mass star (q103q\approx 10^{-3}) significantly heat a disk with α=103\alpha=10^{-3} (see their figure 6).

It is worth pointing out that the secondary outer shock produces a bump in Qsh/QvisQ_{\rm sh}/Q_{\rm vis} at r2rpr\approx 2r_{\rm p}. This bump is large enough for QshQ_{\rm sh} to exceed QvisQ_{\rm vis} in the case of q=103q=10^{-3} and α=104\alpha=10^{-4}. This suggests that the secondary arms may have a non-negligible impact on the disk’s thermal structure away from the planetary orbit. This effect warrants further investigation in future hydrodynamical simulations with wider radial computational domains.

4.2 Caveats and future work

We have demonstrated that shock heating by planet-induced spiral arms may serve as an important heating source for the disk harboring the planet. However, our simulations have two important limitations that should be addressed in future work. Firstly, the computational domain employed in our simulations only covers r>0.5rpr>0.5r_{\rm p} due to computational cost. As a result, we could not fully observe how the secondary inner arm develops into a shock. Meanwhile, we have seen that the heating by the fully developed secondary outer shock can dominate over that by the primary outer shock, when compared at the same orbital radius. This is likely the case for the inner shocks as well, since we observe that the secondary shock’s δ\delta exceeds that of the primary shock already at the inner boundary of our computational domain (see, e.g., figure 3). Simulations with an extended inner computational domain, down to r0.25rpr\sim 0.25r_{\rm p}, will allow us to confirm this expectation. Previous studies (Zhu et al., 2015; Bae et al., 2017) already showed that the tertiary inner arm can also emerge at r0.5rpr\lesssim 0.5r_{\rm p}. Shock heating by this tertiary arm may potentially dominate over that by the secondary arm at large distances from the planet.

Secondly, our simulation models are 2D and therefore neglect any effects arising from the vertical structure of the disk and spiral arms. The spiral shocks can be weaker in 3D disks, where the shocked gas is allowed to expand vertically, leading to adiabatic cooling (Bate et al., 2003; Boley & Durisen, 2006; Lyra et al., 2016). Therefore, our 2D simulations could overestimate the magnitude of δ\delta. Nevertheless, it is possible that the scaling of δ\delta with the model parameters and rr found in this study also holds in 3D disks. If this is the case, our 2D approximation would primarily affect the numerical prefactors in our empirical formulas for δ\delta, equations (20) and (21). In any case, 3D simulations will be necessary to validate our empirical formulas quantitatively.

The simple β\beta-relaxation prescription employed in this study may affect our disk’s thermal structure. However, we expect that this treatment primarily influences disk cooling and does not significantly impact the shock heating rate. In fact, our simulations show that varying the dimensionless thermal relaxation timescale β\beta over the range 10β102.510\leq\beta\leq 10^{2.5} has little effect on δ\delta (see figure 10), indicating that the shock heating strength is insensitive to the thermal relaxation process as long as the relaxation timescale is much longer than the local orbital timescale. For smaller β\beta (0.01\sim 0.0111), radiative damping of planet-induced spiral shocks (Miranda & Rafikov, 2020) could affect δ\delta profiles. However, we have not explored this parameter range as shock heating would have little impact on disk temperature in such rapidly cooling disks.

5 Conclusion

We performed 2D viscous hydrodynamical simulations of protoplanetary disks with an embedded planet to investigate disk heating caused by the shocks associated with planet-induced spiral density waves. We quantified the shock heating rate produced by the primary and secondary spiral arms by directly measuring the specific entropy jumps, δ\delta, across the shocks. We found that the radial profiles of δ\delta approximately follow universal scaling relations that depend only on the planet-to-star mass ratio qq, the dimensionless disk viscosity α\alpha, and the normalized orbital radius (figures 11 and 12). Notably, δ\delta is independent of the radial distributions of the disk’s surface density, sound velocity, and thermal relaxation timescale, as long as the entropy jumps are measurable. We provided analytic expressions for δ\delta for the primary inner and outer arms (equation (20)), as well as for the secondary outer arm (equation (21)). These expressions are accurate within a factor of two for a moderately viscous (103.5α10310^{-3.5}\lesssim\alpha\lesssim 10^{-3}) and moderately adiabatic (10β102.510\lesssim\beta\lesssim 10^{2.5}) disk with a planet massive enough that its spiral arms are strongly nonlinear (qqthq\gtrsim q_{\rm th}). For cases with lower viscosity or a smaller planet, the expressions are less accurate, with deviations up to a factor of three for q=104.5q=10^{-4.5} and α=104\alpha=10^{-4} (RUNs 6 and 8, respectively).

The empirically obtained expressions for the shock heating rate enables one to predict the shock and viscous heating rates for various values of qq and α\alpha. We found that shock heating can dominate over viscous heating when α103\alpha\lesssim 10^{-3} and q103q\gtrsim 10^{-3} (figure 13). Our formulas can also be easily implemented into radially one-dimensional models of gas and dust evolution (e.g., Oka et al., 2011; Bitsch et al., 2015; Delage et al., 2023) and will thus enable the study of how planet-induced shock heating affects gas and dust evolution, as well as planet formation, after a massive planet has formed.

Our results are based on 2D simulations with a relatively narrow inner computational domain. Simulations with an extended radial domain are necessary to model shock heating by the secondary and/or tertiary inner arms. Additionally, 3D simulations are needed to quantify the effects of vertical dynamics on the shock heating rate.

{ack}

We thank Hidekazu Tanaka for useful discussions, Ryosuke Tominaga for useful comments on the treatment of viscous heating in the Athena++ code, and the anonymous referee for insightful and constructive comments that helped to improve the quality of this work.

Funding

This work was supported by JSPS KAKENHI Grant Numbers JP20H00182, JP23H01227, JP23K25923, and JP23K03463.

Data availability

The data underlying this article will be shared on reasonable request to the corresponding author.

References

  • Andrews (2020) Andrews, S. M. 2020, ARA&A, 58, 483–528
  • Bae et al. (2023) Bae, J., Isella, A., Zhu, Z., et al. 2023, in Astronomical Society of the Pacific Conference Series, Vol. 534, Astronomical Society of the Pacific Conference Series, ed. S. Inutsuka, Y. Aikawa, T. Muto, K. Tomida, & M. Tamura, 423
  • Bae & Zhu (2018) Bae, J., & Zhu, Z. 2018, ApJ, 859, 118
  • Bae et al. (2017) Bae, J., Zhu, Z., & Hartmann, L. 2017, ApJ, 850, 201
  • Bate et al. (2003) Bate, M. R., Lubow, S. H., Ogilvie, G. I., & Miller, K. A. 2003, MNRAS, 341, 213
  • Bitsch et al. (2015) Bitsch, B., Johansen, A., Lambrechts, M., & Morbidelli, A. 2015, A&A, 575, A28
  • Bitsch et al. (2019) Bitsch, B., Raymond, S. N., & Izidoro, A. 2019, A&A, 624, A109
  • Boley & Durisen (2006) Boley, A. C., & Durisen, R. H. 2006, ApJ, 641, 534
  • Brown & Ogilvie (2024) Brown, J. J., & Ogilvie, G. I. 2024, MNRAS, 534, 39
  • Delage et al. (2023) Delage, T. N., Gárate, M., Okuzumi, S., et al. 2023, A&A, 674, A190
  • Fukuhara & Okuzumi (2024) Fukuhara, Y., & Okuzumi, S. 2024, PASJ, 76, 708
  • Fung & Dong (2015) Fung, J., & Dong, R. 2015, ApJ, 815, L21
  • Gammie (2001) Gammie, C. F. 2001, ApJ, 553, 174
  • Goldreich & Tremaine (1979) Goldreich, P., & Tremaine, S. 1979, ApJ, 233, 857
  • Goodman & Rafikov (2001) Goodman, J., & Rafikov, R. R. 2001, ApJ, 552, 793
  • Johansen et al. (2021) Johansen, A., Ronnet, T., Bizzarro, M., et al. 2021, Science Advances, 7, eabc0444
  • Kleine et al. (2020) Kleine, T., Budde, G., Burkhardt, C., et al. 2020, Space Sci. Rev., 216, 55
  • Kondo et al. (2023) Kondo, K., Okuzumi, S., & Mori, S. 2023, ApJ, 949, 119
  • Kruijer et al. (2020) Kruijer, T. S., Kleine, T., & Borg, L. E. 2020, Nature Astronomy, 4, 32
  • Lichtenberg et al. (2019) Lichtenberg, T., Golabek, G. J., Burn, R., et al. 2019, Nature Astronomy, 3, 307
  • Lin & Papaloizou (1993) Lin, D. N. C., & Papaloizou, J. C. B. 1993, in Protostars and Planets III, ed. E. H. Levy & J. I. Lunine, 749
  • Lynden-Bell & Pringle (1974) Lynden-Bell, D., & Pringle, J. E. 1974, MNRAS, 168, 603
  • Lyra et al. (2016) Lyra, W., Richert, A. J. W., Boley, A., et al. 2016, ApJ, 817, 102
  • Masset (2000) Masset, F. 2000, A&AS, 141, 165
  • Miranda & Rafikov (2020) Miranda, R., & Rafikov, R. R. 2020, ApJ, 892, 65
  • Mori et al. (2021) Mori, S., Okuzumi, S., Kunitomo, M., & Bai, X.-N. 2021, ApJ, 916, 72
  • Müller et al. (2012) Müller, T. W. A., Kley, W., & Meru, F. 2012, A&A, 541, A123
  • Oka et al. (2011) Oka, A., Nakamoto, T., & Ida, S. 2011, ApJ, 738, 141
  • Pringle (1981) Pringle, J. E. 1981, ARA&A, 19, 137
  • Rafikov (2002) Rafikov, R. R. 2002, ApJ, 569, 997
  • Rafikov (2016) Rafikov, R. R. 2016, ApJ, 831, 122
  • Richert et al. (2015) Richert, A. J. W., Lyra, W., Boley, A., Mac Low, M.-M., & Turner, N. 2015, ApJ, 804, 95
  • Sato et al. (2016) Sato, T., Okuzumi, S., & Ida, S. 2016, A&A, 589, A15
  • Shakura & Sunyaev (1973) Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 500, 33
  • Stone et al. (2020) Stone, J. M., Tomida, K., White, C. J., & Felker, K. G. 2020, ApJS, 249, 4
  • Val-Borro et al. (2006) Val-Borro, M. D., Edgar, R. G., Artymowicz, P., et al. 2006, MNRAS, 370, 529
  • Zhu et al. (2015) Zhu, Z., Dong, R., Stone, J. M., & Rafikov, R. R. 2015, ApJ, 813, 88
  • Ziampras et al. (2020) Ziampras, A., Ataiee, S., Kley, W., Dullemond, C. P., & Baruteau, C. 2020, A&A, 633, A29