Brought to you by:

Evolution of a Mode of Oscillation within Turbulent Accretion Disks

and

Published 2021 October 21 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Robert V. Wagoner and Celia R. Tandon 2021 ApJ 920 114 DOI 10.3847/1538-4357/ac1868

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/920/2/114

Abstract

We investigate the effects of subsonic turbulence on a normal mode of oscillation (a possible origin of the high-frequency quasi-periodic oscillations (HFQPOs) within some black hole accretion disks). We consider perturbations of a time-dependent background (steady-state disk plus turbulence), obtaining an oscillator equation with stochastic damping, (mildly) nonlinear restoring, and stochastic driving forces. The (long-term) mean values of our turbulent functions vanish. In particular, turbulence does not damp the oscillation modes, so "turbulent viscosity" is not operative. However, the frequency components of the turbulent driving force near that of the mode can produce significant changes in the amplitude of the mode. Even with an additional (phenomenological constant) source of damping, this leads to an eventual "blowout" (onset of effects of nonlinearity) if the turbulence is sufficiently strong or the damping constant is sufficiently small. The infrequent large increases in the energy of the mode could be related to the observed low duty cycles of the HFQPOs. The width of the peak in the power spectral density (PSD) is proportional to the amount of nonlinearity. A comparison with observed continuum PSDs indicates the conditions required for the visibility of the mode.

Export citation and abstract BibTeX RIS

1. Introduction

Consider ideal Newtonian hydrodynamics (Thorne & Blandford 2017). This is a useful first approximation in the following exploratory analysis of the interaction of a normal mode of oscillation with turbulence. Although the initial application will be to (geometrically thin) black hole accretion disks, the effects of general relativity should not change the general nature of our results (due to the Principle of Equivalence, within the small volume of the mode). The effects of the magnetic fields within such disks on the mode are unclear and will be discussed in the final section.

The changes that we find in the energy of the mode are produced mainly by the turbulent driving force. This amplification may be relevant to the high-frequency quasi-periodic oscillations (HFQPOs) observed in some of the stellar-mass to supermassive black hole (Figure 1) sources (Smith et al. 2018, 2021). They typically have stable frequencies but low duty cycles. The opposite is the case for the black hole low-frequency QPOs and most of the QPOs of accreting neutron stars (Remillard & McClintock 2006).

Figure 1.

Figure 1. An X-ray light curve (top) and power spectral density (right, showing the HFQPO with a period of 1.8 hr) of the narrow-line Seyfert 1 active galactic nucleus Mrk 766, powered by a black hole of mass 7 × 106 solar masses (Zhang et al. 2017).

Standard image High-resolution image

Our major focus is the evolution of the amplitude of the mode. Previous investigations of oscillators subject to stochastic damping (Grue & Øksendal 1997; Ortega-Rodríguez et al. 2020) have mainly considered time averages and took the damping to have a positive mean value.

We take the background model of the accretion disk to be composed of two components: (a) the stationary and axisymmetric thin disk, and (b) the contribution of subsonic turbulence (continuously generated by the strong magnetorotational instability, which transfers the large free energy of the rate of shear (rdΩ/dr) in the disk into the turbulent cascades (Beckwith et al. 2011)). Thus, the background velocity, pressure, and mass density are decomposed as

Equation (1)

with ∣ ∇ · v ∣ ≪ ∣ ∇ × v ∣ and the turbulent Mach number ${ \mathcal M }\equiv v/{c}_{s}\lt 1$, where cs is the speed of sound.

We consider an adiabatic perturbation of this background from a single normal mode and ignore mode−mode interactions. We also assume that the turbulence is unaffected by the mode. Then, the displacement vector of the mode is taken to be of the form

Equation (2)

where the eigenfunction $\hat{{\boldsymbol{\xi }}}$ (taken to be axisymmetric, a property of many of the most observable modes) is approximated as that calculated from the zero-order steady background model, and the amplitude A is dimensionless. The density fluctuation produced by the mode is Δρ/ρ ≅ −∇ · ξ A. The perturbation theory that we shall now employ therefore requires that A2 ≲ 1.

2. Evolution of the Mode

To analyze the dynamics of the mode, we employ the approach of Schenk et al. (2001) (especially Appendix I). With d/dt = ∂/∂t + u · ∇, and allowing for the time dependence of the unperturbed background model, Equations (I24) and (I25) of Schenk et al. (2001) give the components

Equation (3)

of the acceleration of the perturbation. We invoke the Cowling approximation (neglecting perturbations of the gravitational potential Φ, because the mass of the disk is much less than that of the black hole) and keep the first- and second-order perturbations of the pressure gradient restoring force. In addition, we add to the equation of motion (I37 of Schenk et al. 2001) the main external driving force per unit volume, the divergence of the turbulent Reynolds stress tensor.

This then gives the full equation of motion

Equation (4)

where Γ1 is the adiabatic index of the perturbed pressure. The quantity ${Q}_{j}^{i}={k}_{1}{{\rm{\Theta }}}_{j}^{i}+{k}_{2}{\rm{\Theta }}{\delta }_{j}^{i}+{k}_{3}{{\rm{\Xi }}}_{j}^{i}+{k}_{4}{\rm{\Xi }}{\delta }_{j}^{i}$, where ∣kn ∣ ∼1, with ${{\rm{\Theta }}}_{j}^{i}={{\rm{\nabla }}}_{k}{\xi }^{k}{{\rm{\nabla }}}_{j}{\xi }^{i}$ and ${{\rm{\Xi }}}_{j}^{i}={{\rm{\nabla }}}_{k}{\xi }^{i}{{\rm{\nabla }}}_{j}{\xi }^{k}$ .

Operating with $\int {d}^{3}x\hat{{\xi }^{i}}$ on the above equation then gives

Equation (5)

with the time-dependent functions generated by the turbulence. (Interactions with other modes would also contribute to the term g1(t), proportional to their amplitudes.) The constant ${c}_{0}=\int {d}^{3}x\rho {\hat{\xi }}^{i}{\hat{\xi }}_{i}$ and the constant ${g}_{0}={c}_{0}{\omega }_{0}^{2}$, where ω0(∼ Ω) is the eigenfrequency of the mode in the absence of turbulence and the nonlinearity in the restoring force. Employing the phase τω0 t and dividing the above equation by c0 gives our master equation

Equation (6)

in which all quantities are dimensionless. From Equation (4), it is seen that $| {N}_{0}| \sim | {{\rm{\nabla }}}_{i}{\hat{\xi }}^{j}| \sim 1$.

Multiplying Equation (6) by dA/d τ gives

Equation (7)

We have introduced the (dimensionless) energy

Equation (8)

of the mode.

We find that

Equation (9)

To obtain the final expression, we have employed an integration by parts and the conservation of mass (∂ρ/∂t + ∇ · (ρ u ) = 0). We see that F1 is proportional to the rate of change of the mass within the mode, so a decreasing mass produces growth of the mode. However, its long time average 〈F1〉 = 0. Therefore, we now include a phenomenological source of damping, so that

Equation (10)

with the constant D* > 0.

The function g1(t) is generated by the last two terms in Equation (3) and the first three terms on the right-hand side of Equation (4) (which are larger than the fourth). The function n1(t) is generated by the term in Equation (4) involving Qi j . Therefore,

Equation (11)

Finally, we obtain

Equation (12)

where ${{ \mathcal F }}_{i}=-{{\rm{\nabla }}}^{j}({\rho }_{0}{v}_{i}{v}_{j})$ is the turbulent driving force per unit volume.

Consider briefly the case when A2 ≪ 1, so that we can neglect the nonlinear term in Equation (6). Employing the change of variable $A(\tau )=a(\tau )\exp [-0.5{\int }_{{\tau }_{0}}^{\tau }{F}_{1}(s){ds}]$ and the upper limits on the magnitudes of the stochastic functions (discussed in the next section) when ${ \mathcal M }\ll 1$ (giving $| {G}_{1}+(1/2){{dF}}_{1}/d\tau +(1/4){F}_{1}^{2}| \ll 1$), we obtain ${d}^{2}a/d{\tau }^{2}+a={F}_{2}(\tau )\exp [0.5{\int }_{{\tau }_{0}}^{\tau }{F}_{1}(s){ds}]$ from Equation (6). Employing the relevant Green's function, we then obtain the solution

Equation (13)

corresponding to the initial conditions A = A0 and ${dA}/d\tau \,={\dot{A}}_{0}$, at τ = τ0 = 0.

Let us consider the evolution of A(τ) at late times (〈F1τ ≫ 1), with $\langle {F}_{1}\rangle \equiv {(\tau -\tau ^{\prime} )}^{-1}{\int }_{\tau ^{\prime} }^{\tau }{F}_{1}(s){ds}\cong {D}_{* }\ll 1$. From Equation (13) we then obtain (until effects of nonlinearity become important)

Equation (14)

where τ* ≡ 1/D* and ττ* ≫ 1.

Including the mild nonlinearity, the evolution of the energy E(τ) of the mode is governed by Equation (7). From Equation (8), the potential energy is PE = A2/2 + N0 A3/3. ∣A∣ becomes unbounded if the energy reaches $1/(6{N}_{0}^{2})$, which occurs when A = − 1/N0, as shown in Figure 2. Will equilibrium be achieved before this blowout can occur? (Of course, "blowout" only indicates that nonlinearities have become important, so the evolution cannot be accurately continued. The energy would remain bounded if the coefficient of an A4 term was positive.) Our observational predictions (in Section 4) are all based upon choices of parameters that do not produce a blowout.

Figure 2.

Figure 2. The potential energy PE = (1/2)A2 + (1/3)N0 A3, illustrating the condition for a "blowout" of the kinetic energy (KE) (onset of the effects of nonlinearity).

Standard image High-resolution image

Averaging over a time interval δ τ (with 2πδ ττ), Equation (7) gives

Equation (15)

(We have found that the contributions of G1 and N1 are negligible). At blowout, $\langle {({dA}/d\tau )}^{2}\rangle \cong \langle {A}^{2}\rangle \cong 1/(2{N}_{0}^{2})$. Initially, the driving term will be larger than the damping term, because it is only proportional to A. Therefore, the condition for equilibrium to be achieved before a blowout could occur is

Equation (16)

3. Modeling the Turbulence

We next consider how to characterize the turbulence. The ensemble and time averages of our four stochastic functions (H, F2, G1, N1) should vanish. The major properties of a turbulent eddy of radius are its velocity v and turnover time t ∼ 2π /v . The smallest dimension (radius) of the largest turbulent eddy will be of order the (half) thickness h(r) of the accretion disk. Vertical force balance gives hΩ ∼ cs .

We shall consider a normal mode with a similar thickness (such as the fundamental g-mode (also called r-mode)), with period tm ∼ 2π/Ω ∼ tL (the period of the largest eddy). Goldreich & Kumar (1988) found that an acoustic mode in the Sun couples most strongly to the eddies of the convective turbulence with t tm , giving ${\ell }\lesssim { \mathcal M }h$ , where ${ \mathcal M }$ is now the Mach number of the largest eddies (radius Lh). We shall assume that the same holds true for the modes in our more strongly turbulent accretion disks (Nowak & Wagoner 1995).

The pressure and density fluctuations produced by such eddies are

Equation (17)

We assume that these fluctuations arise mainly from the conversion of the eddies into acoustic modes (Goldreich & Kumar 1988) and take ∣ ∇ ∣ ≲ 1/h. A limit on the turbulent velocity comes from the effects of a "turbulent viscosity" on the structure and evolution of accretion disks (Kato et al. 2008). Beckwith et al. (2011) find that ${ \mathcal M }\approx 0.06\mbox{--}0.10$.

We generate the (now dimensionless) eddy lifetimes Δτ (assumed equal to the turnover time τ ) by random sampling a distribution function with mean 〈Δτ〉 ≲ π, and we shall assume no delay between eddies. (Note that the unperturbed period of the mode is τm = 2π.) We choose a Rayleigh distribution $f{({\rm{\Delta }}\tau )=({\rm{\Delta }}\tau /{\sigma }_{* }^{2})\exp [-({\rm{\Delta }}\tau )}^{2}/(2{\sigma }_{* }^{2})]$. Its mean $\langle {\rm{\Delta }}\tau \rangle \,={\sigma }_{* }\sqrt{\pi /2}$, and its variance is $(2-\pi /2){\sigma }_{* }^{2}$.

From the above equations, it is seen that

Equation (18)

where η is the ratio of the volume of the dominant correlated eddies within the mode to the volume of the mode. Thus, because $\eta {{ \mathcal M }}^{2}\ll 1$, we neglect G1 and N1 (compared to unity) in our master Equation (6). Also, if F1 is given by Equation (10), we again see that H(τ) does not affect the long-term solution (Equation (13)) to our linear equation, because ${\int }_{{\tau }_{1}}^{{\tau }_{2}}{F}_{1}(s){ds}\,=[H({\tau }_{2})-H({\tau }_{1})]+{D}_{* }({\tau }_{2}-{\tau }_{1})\cong {D}_{* }({\tau }_{2}-{\tau }_{1})$.

We approximate the functions H(τ) and F2(τ) as

Equation (19)

during each eddy lifetime Δτn . Note that ∣Yk,n ∣ is the maximum value of ∣Ψk ∣ in eddy n, and the value and first derivative of Ψk vanish at the beginning and end of the eddy.

The values of Yk,n are generated by random sampling a Gaussian–Markov conditional probability function $P({Y}_{k,n},{\bar{\tau }}_{n}| {Y}_{k,n-1})$, where ${\bar{\tau }}_{n}=({\rm{\Delta }}{\tau }_{n}+{\rm{\Delta }}{\tau }_{n-1})/2$. Its mean $\langle {Y}_{k,n}\rangle =\bar{Y}+{e}^{-{\bar{\tau }}_{n}/{\tau }_{r}}({Y}_{k,n-1}-\bar{Y})$, and its variance ${\sigma }_{k,n}^{2}\,=(1-{e}^{-2{\bar{\tau }}_{n}/{\tau }_{r}}){\sigma }_{k}^{2}$. We choose the equilibrium mean $\bar{Y}=0$, consistent with our assumption that 〈Ψk 〉 = 0 for averaging times δ τ ≫ Δτ. We choose the relaxation time τr = K〈Δτ〉, with the value of K allowing for correlations between subsequent eddies. We usually choose values of ${\sigma }_{k}=\eta {{ \mathcal M }}^{2}\lesssim {10}^{-3}$, consistent with the expectation that η ≪ 1 and ${{ \mathcal M }}^{2}\lesssim {10}^{-2}$.

4. Results

4.1. Evolutions

For a calculation characterized by particular values of the physical parameters ($\eta {{ \mathcal M }}^{2}$, D*, 〈Δτ〉, and K), we employ a second-order integrator, Huen's method. We generate M samples of Δτ and Ψk (τ). Each evolution has a duration ${\tau }_{\max }\approx M\langle {\rm{\Delta }}\tau \rangle $ , with different random realizations of the distributions of Δτ and Yk during each eddy n. The initial conditions are chosen to be ${A}_{0}={\dot{A}}_{0}=0$ at τ = 0, but the long-term evolution does not depend on them.

In Figure 3, we see how short-term changes in 〈F2 dA/d τδ τ correlate with changes in the energy E. The damping function 〈F1(dA/d τ)2δ τ is usually subdominant on this averaging timescale, but notice that antidamping becomes strong just before the blowout. As predicted in Section 2, the energy reaches E = 1/6 at blowout for our choice N0 = 1.

Figure 3.

Figure 3. An evolution of the energy E of the mode and that of short-term (δ τ = 104) averages of the two stochastic functions (F2(dA/d τ) and F1(dA/d τ)2), which mainly determine it, for $\eta {{ \mathcal M }}^{2}={10}^{-3}$, D* = 10−6, 〈Δτ〉 = π, and K = 1. A blowout occurs at τ = 4.4 × 106.

Standard image High-resolution image

In Figure 4, we see how the negative extent of the amplitude A grows as the blowout is approached and reaches A = −1/N0 = −1 at blowout. It is also seen that the period of the mode has increased by a factor of about 1.7 since the beginning of the evolution, as also expected from Section 2. Also note the behavior of the driving and damping functions during the eddies.

Figure 4.

Figure 4. A short (M = 102) segment at the end of the evolution of A, from the same run as shown in Figure 3, now including F2 and F1.

Standard image High-resolution image

In Figure 5, the damping constant has been increased by a factor of 10, which is sufficient to prevent a blowout (at least over the extended time interval indicated). This is our fiducial evolution. The maximum value of the energy (about 1/2 that required for a blowout) is about 10 times larger than its average (〈E〉 = 8 × 10−3 ).

Figure 5.

Figure 5. Same as Figure 3, but for D* = 10−5. No blowout occurs, at least before the end of this run at τ = 2 × 107.

Standard image High-resolution image

In Figure 6, the parameter choices that produce a blowout or not are indicated, with the corresponding rms amplitudes. Consider times ττ* = 1/D*, so we can employ Equation (14) while the effects of nonlinearity are small. Then we obtain from Equation (15), when averaging over times δ τ = 104,

Equation (20)

where $S(\tau )={F}_{2}(\tau ){\int }_{0}^{{\tau }_{* }}[{{dF}}_{2}(\tau -x)/d\tau ]\sin (x){dx}$. As $S\,\propto {(\eta {{ \mathcal M }}^{2})}^{2}$ and the integral spans τ* = 1/D*, $\langle E\rangle (\tau )\,\propto {(\eta {{ \mathcal M }}^{2})}^{2}/{D}_{* }$ . Because the blowout occurs at a fixed value of E = 1/6, the boundary of the blowout region does appear to approximately agree with the dependence ${D}_{* }\sim 10{(\eta {{ \mathcal M }}^{2})}^{2}$.

Figure 6.

Figure 6. The dependence of the rms amplitude of the mode on the driving and damping parameters.

Standard image High-resolution image

From Figure 7, we see how 〈A2〉 and $\langle {F}_{2}^{2}{\rangle }^{1/2}$ depend on our four parameters. The dependence $\langle {A}^{2}\rangle \approxeq \langle E\rangle \sim 0.1{D}_{* }^{-1}{(\eta {{ \mathcal M }}^{2})}^{2}$ is also obtained from this figure.

Figure 7.

Figure 7. The dependence of the values of 〈A2〉 and $\langle {F}_{2}^{2}{\rangle }^{1/2}$ on the parameters D*, $\eta {{ \mathcal M }}^{2}$, K, and 〈Δτ〉. Averages are over the entire evolution. In each case, the other parameters are at their fiducial value, indicated by their central values on the bottom axes.

Standard image High-resolution image

How does the average energy in the mode compare to the average energy in the effective turbulent eddies? The energy in the mode is approximately ∫d3 x ρ(∂ξi /∂t)(∂ξi /∂t) ≈ c0 ω2 A2. The energy in a dominant eddy is approximately 0.5∫d3 x ρ v2c0 ω2 F2. Thus, equipartition would imply that $\langle {A}^{2}\rangle \approx \langle {F}_{2}^{2}{\rangle }^{1/2}$. The results shown in Figure 7 indicate that this is approximately true for small values of $\eta {{ \mathcal M }}^{2}$.

4.2. Contribution of the Mode to an Observed Power Spectral Density

Consider an axisymmetric (the most observable) mode of oscillation at radius rm with volume v = 2π rm ζ h2(rm ). Its observed normalized power spectral density (PSD) Px (f) (f = ω/2π) is related to the the photon count rate x by

Equation (21)

where Δf is its FWHM. Adopting the approach of Nowak & Wagoner (1995),

Equation (22)

at r = rm . The average photon number flux from the disk is ${ \mathcal F }(r)={\int }_{{\nu }_{1}}^{{\nu }_{2}}{{ \mathcal F }}_{\nu }(\nu ,r)\xi (\nu )d\nu $ for a detector efficiency ξ(ν). We shall employ the approximation ${ \mathcal F }{(r)={ \mathcal F }({r}_{m})(r/{r}_{m})}^{-{C}_{1}}$.

In addition, recall that the density fluctuation Δρ/ρA(τ), related to its (not normalized) PSD by

Equation (23)

The average count rate is (assuming that C1 > 2)

Equation (24)

giving

Equation (25)

Now consider the observed continuum PSD Px (cont.) of black hole sources near the frequency f0 of an HFQPO, with (M/10M)f0 ∼ 200 Hz for black hole binaries (BHBs) and ∼50 Hz for NLS1 AGNs (Smith et al. 2021). We find that f0 Px (cont.) ∼ (2–20) × 10−4 for BHBs and ∼ (3 − 30) × 10−3 for NLS1s. The corresponding quantity for our mode is given (from Equation (25)) by

Equation (26)

The quality factor Qf0f and $\chi \sim {\left(\zeta h/r\right)}^{2}$ .

During a quasi-equilibrium, when E(τ) is changing slowly but is not far below its blowout value of 1/6 (for ${N}_{0}^{2}=1$), the effects of our lowest-order nonlinearity control the width of the PSD peak, with the damping and driving forces subdominant. Then the standard analysis (Landau & Lifshitz 1976) predicts an amplitude-dependent shift of the anharmonic oscillator frequency of

Equation (27)

If the energy varies over a range ΔE during an evolution, it produces a corresponding width Δq of the PSD peak. Comparing the range of energies (that occupy a significant fraction of of the time) seen in Figure 5 with the width of the PSD peak corresponding to the same (fiducial) parameters in Figures 8 and 9, we see that this relation is approximately valid. It is only indicative, as the effects of higher-order nonlinearities have been neglected. From the results shown in Figure 9, we obtain the approximate relation Δq ∼ 1.0〈A20.9.

However, it is also seen from Figures 8 and 9 (and their extension for other values of the parameters) that the maximum value of each of the PSDs of the fundamental mode is P0 ∼ 3, where ∫PA (q)dq ≈ 1.1P0Δq. Therefore, employing Equations (23) and (26), we see that visibility of an HFQPO (Px (max.,QPO) > Px (cont.)) requires that

Equation (28)

Recalling that χ ∼ (ζ h/r)2, typical values of ζ ∼ 1 and h/r ∼ 10−2 and the requirement A2 < 1 could produce a visible HFQPO if Q ≫ 1 or χχ(typical) ∼ 10−4. For instance, O'Neill et al. (2009) found that the trapped g-mode leaked into outgoing p-waves in their viscous hydrodynamic simulation. This would increase the value of ζ = Δr/h.

Figure 8.

Figure 8. Power spectral density of the amplitude A of the mode, with qω/ω0. The two smallest values of the damping parameter that did not produce a blowout (for the fiducial value of $\eta {{ \mathcal M }}^{2}$) are chosen. The harmonics (q = 1, 2, 3) of the unperturbed linear oscillator are indicated by the red lines.

Standard image High-resolution image
Figure 9.

Figure 9. Detail of the fundamental mode shown in the previous figure, with a Gaussian fit. Then ∫PA (q)dq ≈ 1.1P0Δq, where P0 is the maximum value of the PSD and Δq is the full width at half maximum.

Standard image High-resolution image

Nowak & Wagoner (1995) estimated the contribution of the turbulence to the continuum PSD. Referring to their Figure 1, it is seen that near the frequencies of the HFQPOs in BHBs,

Equation (29)

Then with LLEdd and ${ \mathcal M }\lesssim 0.1$, we obtain f0 Px (turb.) ≲ 10−5. Comparing with the observed values of f0 Px (cont.) above, we see that it is unlikely that turbulence is the major contributor to the observed continuum PSD. This result is consistent with the fact that at lower frequencies, it is also seen from their Figure 1 that the turbulence in the accretion disk at the correspondingly larger radii provides very little of the observed power in the X-ray band.

5. Discussion

A critical feature of the physical conditions that we have investigated is the fundamental difference between the nature of our turbulent damping function dH/d τ and the "turbulent viscosity" νt v /3 (analogous to molecular viscosity) employed in the analysis of the effects of turbulence acting on the rate of shear of a quasi-steady flow (Thorne & Blandford 2017). See in particular Kato et al. (2008, Section 11.5). Because the contribution of turbulence to F1(τ) is a total time derivative, there is an equal probability of short-term positive and negative damping. There should be no long-term temporal correlations between the mode and the turbulent eddies.

The physical origin of the damping of the p-modes in the Sun is uncertain (Basu 2016), but it should involve the coupling to higher frequency p-waves, which are damped when their wavelength becomes less than the scale height at the photosphere. This may also be true for our accretion disks. The generation of Alfvén waves could also contribute to the damping. Nowak et al. (1997) found that changes in entropy produced by radiative transfer effects led to the growth of modes within the types of accretion disks considered here.

We hope to consider magnetic forces in the future. In relevant numerical MHD simulations, although the ratio of magnetic to gas pressure is typically a few percent, the ratio of magnetic pressure to the Reynolds stress can be greater than unity (Dewberry et al. 2020). In addition, the magnetic forces can reduce the trapping of the g-modes (Fu & Lai 2009), but the amount depends on the relative magnitude of the poloidal and toroidal components (Ortega-Rodríguez et al. 2015). Reynolds & Miller (2009) found that the g-mode did not appear in their GRMHD simulations, although the number of orbits may have been insufficient to see growth of the mode. However, Dewberry et al. (2020) found that a small eccentricity in accretion disk orbits can excite r (g)-modes to large amplitudes in their MHD simulations. Warps can also excite such modes (Kato 2004, 2008).

In the future, we also hope to refine the modeling of the turbulence. One issue is the collective correlated effect of the eddies within the mode. Another is the effects of anisotropy, in particular the stretching of eddies in the e ϕ direction. In addition, could (a) the infrequent significant increases of the mode energy or (b) a temporary reduction in the damping or (c) intermittency in the turbulent cascade be relevant to the low duty cycle exhibited by the HFQPOs?

The approach that we have taken to this problem may be applicable to other physical systems in which an oscillator is coupled to a turbulent environment.

We thank Nicole Lloyd-Ronning, Jeff Scargle, and James Stone for helpful comments. C.R.T. acknowledges financial support from the Stanford Physics Department Summer Undergraduate Research Program and NASA award 80NSSC20K0591 to Krista Lynne Smith. We would like to thank Stanford University and the Stanford Research Computing Center for providing computational resources on the Sherlock cluster and support. Support was also provided by the Center for Space Science and Astrophysics at Stanford. We thank the referee for helpful comments.

Please wait… references are loading.
10.3847/1538-4357/ac1868