1 Introduction

There is now ample evidence for the existence of low-frequency waves and oscillations in the structured solar atmosphere (for recent reviews, see e.g., Nakariakov and Verwichte, 2005; Banerjee et al., 2007; Roberts, 2008; De Moortel and Nakariakov, 2012; Liu and Ofman, 2014). When combined with magnetohydrodynamic (MHD) theory, the measured wave parameters allow inferring the solar atmospheric parameters that are difficult to measure directly. This practice was originally proposed for the solar corona (Roberts, Edwin, and Benz, 1984; see also Rosenberg, 1970; Uchida, 1970; Zajtsev and Stepanov, 1975), but has been extended to spicules (e.g., Zaqarashvili and Erdélyi, 2009), prominences (e.g., Arregui, Oliver, and Ballester, 2012), magnetic pores (e.g., Morton et al., 2011), and various structures in the chromosphere (e.g., Jess et al., 2009; Morton et al., 2012), to name but a few. Compared with sausage waves (with azimuthal wavenumber \(m=0\)), kink waves (with \(m=1\)) have received more attention, presumably because of their ubiquity in the solar atmosphere (e.g., Nakariakov et al., 1999; Aschwanden et al., 1999; Tomczyk and McIntosh, 2009; Kupriyanova, Melnikov, and Shibasaki, 2013). However, recent observations indicated that sausage waves abound as well (e.g., Nakariakov, Melnikov, and Reznikova, 2003; Morton et al., 2012; Grant et al., 2015; Moreels et al., 2015). In addition, standing sausage modes in flare loops were shown to be important for interpreting a considerable fraction of quasi-periodic pulsations (QPPs) in the light curves of solar flares (see Nakariakov and Melnikov (2009) for a recent review).

A theoretical understanding of fast sausage waves supported by magnetized cylinders is crucial for their seismological applications. For this purpose, the transverse density distribution is usually idealized as being in a step-function (top-hat) form, characterized by the internal (\(\rho_{\mathrm{i}}\)) and external (\(\rho_{\mathrm{e}}\)) values (e.g., Meerson, Sasorov, and Stepanov, 1978; Spruit, 1982; Edwin and Roberts, 1983; Cally, 1986; Kopylova et al., 2007; Vasheghani Farahani et al., 2014). In a low-\(\beta\) environment such as the solar corona, two regimes are known to exist, depending on the longitudinal wavenumber \(k\) (e.g., Spruit, 1982). When \(k\) exceeds some critical \(k_{\mathrm{c}}\), the trapped regime arises, whereby the sausage wave energy is well confined to the cylinder. In contrast, if \(k< k_{\mathrm{c}}\), then the leaky regime results, and fast sausage waves experience apparent temporal damping by emitting their energy into the surrounding fluid. Furthermore, the \(k\)-dependence of the periods \(P\) and damping times \(\tau\) of leaky waves disappears when \(k\) is sufficiently small (e.g., Kopylova et al., 2007; Vasheghani Farahani et al., 2014). Let \(R\) denote the cylinder radius, and \(v_{\mathrm{Ai}}\) denote the internal Alfvén speed. In the long-wavelength limit (\(k\rightarrow0\)), \(P\) is found to be primarily determined by the transverse Alfvén transit time \(R/v_{\mathrm{Ai}}\), while the ratio \(\tau/P\) is largely proportional to the density contrast \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\) (Kopylova et al., 2007). This then enables one to employ the measured \(P\) and \(\tau\) to deduce \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\) and \(R/v_{\mathrm{Ai}}\), with the latter carrying important information on the magnetic field strength in flare loops.

Evidently, there is no reason to expect that the density distribution across magnetic cylinders is in a step-function fashion. This has stimulated a series of studies to examine the properties of fast sausage waves in magnetized cylinders with a continuous transverse density profile by either proceeding analytically with an eigenmode analysis (Edwin and Roberts, 1988; Lopin and Nagorny, 2014, 2015) or numerically solving the linearized MHD equations as an initial-value problem (Nakariakov, Hornsey, and Melnikov, 2012; Chen et al., 2015a). Many features in the step-function case, the \(k\)-dependence in particular, were found to survive. However, the period \(P\) (Nakariakov, Hornsey, and Melnikov, 2012) and damping time \(\tau\) (Chen et al., 2015a) may be sensitive to yet another parameter, namely the steepness or equivalently the length scale of the transverse density inhomogeneity. We note that the steepness is crucial in determining such coronal heating mechanisms as resonant absorption (e.g., Hollweg and Yang, 1988; Goossens, Andries, and Aschwanden, 2002; Ruderman and Roberts, 2002) and phase mixing (Heyvaerts and Priest, 1983). There is then an obvious need to employ the measured \(P\) and \(\tau\) of sausage modes to infer the profile steepness, in much the same way that kink modes were employed (Arregui et al., 2007; Goossens et al., 2008; Soler et al., 2014). This was undertaken by Chen et al. (2015b, hereafter Paper I), based on an analytical dispersion relation (DR) governing linear fast sausage waves in cylinders with a rather general transverse density distribution. The only requirement was that this profile can be decomposed into a uniform cord, a uniform external medium, and a transition layer connecting the two. However, this layer can be of arbitrary width and the profile therein can be in arbitrary form.

The aim of the present study is to extend the analysis in Paper I in the following aspects. First, we remove the restriction for the transverse density profile to involve a uniform cord, thereby enabling the analysis to be applicable to a richer variety of density distributions. Second, when validating the results from this eigenmode analysis, we employ an independent approach by solving the time-dependent version of linear MHD equations. We detail these time-dependent computations pertinent to the afore-mentioned transverse density profile. Third, we extend the seismological applications in Paper I to QPP events that involve both kink and sausage modes. To illustrate the scheme for inverting multi-mode measurements, Paper I adopted the analytical expressions for the kink mode period and damping time in the thin-tube-thin-boundary (TTTB) limit as given by Goossens et al. (2008). In this study we replace the TTTB expressions with a self-consistent, linear, resistive MHD computation. This is necessary given that flare loops tend not to be thin (Aschwanden, Nakariakov, and Melnikov, 2004), and it is not safe to assume a priori that the density inhomogeneity takes place in a thin transition layer. Fourth, in connection with the third point, we take this opportunity to provide a very detailed examination of resonantly damped kink modes in cylinders with transverse density profiles in question.

This manuscript is organized as follows. Section 2 presents the necessary equations, the derivation of the analytic DR in particular. The behavior of sausage waves in nonuniform cylinders and its applications to QPP events are then presented in Section 3. Finally, Section 4 summarizes the present study.

2 Mathematical Formulation

2.1 Derivation of the Dispersion Relation

Appropriate for the solar corona, we adopted ideal, cold (zero-\(\beta\)) MHD to describe fast sausage waves. The magnetic loops hosting these waves were modeled as straight cylinders with radius \(R\) aligned with a uniform magnetic field \({\boldsymbol {B}} = B\hat{{\boldsymbol {z}}}\), where a standard cylindrical coordinate system \((r, \theta, z)\) was adopted. The equilibrium density was assumed to be a function of \(r\) only and of the form

$$ \rho(r)=\left \{ \textstyle\begin{array}{l@{\quad}l} \rho_{\mathrm{i}} [1- (1-\frac{\rho_{\mathrm{e}}}{\rho _{\mathrm{i}}} )f(r) ], & 0 \le r < R,\\ \rho_{\mathrm{e}}, & r > R, \end{array}\displaystyle \right . $$
(1)

where \(f(r)\) is some arbitrary function that increases smoothly from 0 at \(r=0\) to unity when \(r=R\). Furthermore, \(\rho_{\mathrm{i}}\) and \(\rho_{\mathrm{e}}\) denote the densities at the cylinder axis and in the external medium, respectively. The corresponding Alfvén speeds follow from the definition \(v_{\mathrm{Ai}, \mathrm{e}} = B/\sqrt{4\pi\rho_{\mathrm{i}, \mathrm{e}}}\).

It suffices to briefly outline the mathematical approach for establishing the pertinent dispersion relation (DR), since this approach has been detailed in Paper I. To start, we focused on axisymmetric sausage perturbations, and Fourier-analyzed any perturbation \(\delta f(r, z, t)\) as

$$\begin{aligned} \delta f(r, z, t) = \mathrm{Re} \bigl\{ \tilde{f}(r)\exp \bigl[-i (\omega t-k z ) \bigr] \bigr\} . \end{aligned}$$
(2)

It then follows from the linearized, ideal, cold MHD equations that the Fourier amplitudes of the transverse Lagrangian displacement (\(\tilde {\xi}_{r}\)) and Eulerian perturbation of total pressure (\(\tilde{p}_{\mathrm{T}}\)) are governed by Equations (6) and (7) in Paper I, respectively. Now that sausage waves do not resonantly couple to torsional Alfvén waves for the configuration we examine, we employed regular series expansions about \(y\equiv r-R/2 = 0\) to express \(\tilde{\xi}_{r}\) and \(\tilde{p}_{\mathrm{T}}\) in the nonuniform portion of the density distribution. Further requiring that sausage waves do not disturb the cylinder axis (\(\tilde{\xi}_{r} = 0\) at \(r=0\)), and employing the conditions for \(\tilde{\xi}_{r}\) and \(\tilde{p}_{\mathrm{T}}\) to be continuous at the interface \(r=R\), we found that the DR can be expressed as

$$\begin{aligned} & \frac{\frac{\mu_{\mathrm{e}}R H_{0}^{(1)}(\mu_{\mathrm{e}} R)}{H_{1}^{(1)}(\mu_{\mathrm{e}} R)}\tilde{\xi }_{1}(R/2) -\tilde{\xi}_{1}(R/2)-R\tilde{\xi}^{\prime}_{1}(R/2)}{\tilde{\xi}_{1}(-R/2)} \\ & \quad=\frac{\frac{\mu_{\mathrm{e}}R H_{0}^{(1)}(\mu_{\mathrm{e}} R)}{H_{1}^{(1)}(\mu_{\mathrm{e}} R)} \tilde{\xi}_{2}(R/2)-\tilde{\xi }_{2}(R/2)-R\tilde{\xi}^{\prime}_{2}(R/2)}{\tilde{\xi}_{2}(-R/2)}. \end{aligned}$$
(3)

Here \(H_{n}^{(1)}\) denotes the \(n\)-th order Hankel function of the first kind, and \(\mu_{\mathrm{e}}\) is defined by

$$\begin{aligned} \mu_{\mathrm{e}}^{2} = {\frac{\omega^{2}}{v^{2}_{\mathrm{Ae}}}-k^{2}} \quad \biggl(-\frac{\pi}{2} < \arg\mu_{\mathrm{e}} \le\frac{\pi}{2} \biggr). \end{aligned}$$
(4)

Furthermore,

$$ \tilde{\xi}_{1}(y)=\sum^{\infty}_{n=0}a_{n} y^{n}\quad\mbox{and}\quad \tilde{\xi}_{2}(y)=\sum ^{\infty}_{n=0}b_{n} y^{n} $$
(5)

are two linearly independent solutions for \(\tilde{\xi}_{r}\) in the portion \(r< R\). Without loss of generality, we chose

$$\begin{aligned} a_{0}=R,\quad\quad a_{1}=0,\quad\quad b_{0}=0,\quad\quad b_{1} =1 . \end{aligned}$$
(6)

The rest of the coefficients \(a_{n}\) and \(b_{n}\) can be found by replacing \(R\) with \(R/2\) in Equation (11) in Paper I and contain the information on the density distribution. Finally, the prime ′ denotes the derivative of \(\tilde{\xi}_{1, 2}\) with respective to \(y\).

Before proceeding, we note that a series-expansion-based approach was recently adopted by Soler et al. (2013) to treat wave modes in transversally nonuniform cylinders where the azimuthal wavenumber \(m\) is allowed to be arbitrary. A comparison between this approach and ours is detailed in Appendix C, where we show that both approaches yield identical results for trapped sausage modes (\(m=0\)). While our approach seems more appropriate to describe leaky sausage modes, we stress that a singular expansion as employed by Soler et al. (2013) is necessary to treat wave modes with \(m\ne0\).

2.2 Solution Method

Throughout this study, we focus on standing sausage modes by restricting longitudinal wavenumbers (\(k\)) to be real, but allowing angular frequencies (\(\omega\)) to be complex-valued (\(\omega= \omega_{\mathrm{R}}+i\omega_{\mathrm{I}}\)). In addition, we focus on fundamental modes, namely those with \(k=\pi /L\) where \(L\) is the loop length. In practice, we started with prescribing an \(f(r)\), and then solved Equation (3) by truncating the infinite series expansion (Equation (5)) to retain terms with \(n\) up to \(N=101\). Using an even larger \(N\) leads to no appreciable difference. It should be noted that \(\omega\) in units of \(v_{\mathrm{Ai}}/R\) depends only on the combination \([f(r), kR, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\). The corresponding period \(P\) and damping time \(\tau\) follow from the definitions \(P=2\pi/\omega _{\mathrm{R}}\) and \(\tau= 1/|\omega_{\mathrm{I}}|\).

For validation purposes, we also obtained \(\omega\) as a function of \(k\) in a way independent of this eigenmode analysis. This was done by solving the time-dependent equation governing the transverse velocity perturbation \(\delta v_{r} (r, z,t) \) as an initial-value problem. For given combinations of \([f(r), kR, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\), the periods and damping times of sausage modes can be found by analyzing the temporal evolution of the perturbation signals (see Appendix A for details). As we show in Figure 2, the values of \(P\) and \(\tau\) derived from the two independent approaches are in close agreement. However, numerically solving the analytical DR is much less computationally expensive. In addition, the values of \(\tau\) for heavily damped modes can be readily found, whereas the perturbation signals in time-dependent computations decay too rapidly to allow a proper determination of \(\tau\).

3 Numerical Results

It is impossible to exhaust the possible prescriptions for \(f(r)\). We therefore focus on one choice, namely

$$\begin{aligned} f(r)= \biggl(\frac{r}{R} \biggr)^{\mu}, \end{aligned}$$
(7)

where \(\mu\) is positive. The density profiles with a number of different \(\mu\) are shown in Figure 1, where \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\) is chosen to be 100 for illustration purposes. Evidently, the profile becomes increasingly steep as \(\mu\) increases and approaches a step-function form when \(\mu\) approaches infinity. This makes it possible to investigate the effect of profile steepness by examining the \(\mu\)-dependence of the numerical results. In addition, for fundamental modes with \(k=\pi/L\), the dependence on \(kR\) is translated into that on the length-to-radius ratio \(L/R\).

Figure 1
figure 1

Transverse equilibrium density profiles as a function of \(r\) for different steepness parameters \(\mu\) as labeled. Here the density contrast \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\) is chosen to be 100 for illustration purposes.

3.1 Behavior of Sausage Waves in Nonuniform Tubes

Figure 2 presents the dependence on \(L/R\) of the period \(P\) and damping time \(\tau\) for a series of \(\mu\) values as labeled. For illustration purposes, the density ratio \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\) is taken to be 100. The dash-dotted line in Figure 2a represents \(P=2L/v_{\mathrm{Ae}}\), and separates trapped (to its left) from leaky (right) modes. The solid curves represent the results from solving the analytical DR (Equation (3)), whereas the circles represent those obtained with the initial-value-problem approach. A close agreement between the curves and circles is clear, thereby validating the DR.

Figure 2
figure 2

Dependence on length-to-radius ratio \(L/R\) of (a) periods \(P\) and (b) damping times \(\tau\) of fundamental sausage modes. A number of density profiles with different \(\mu\) are examined as labeled. The black dash-dotted line in (a) represents \(P=2L/v_{\mathrm{Ae}}\) and separates the trapped (to its left) from leaky (right) regimes. The open circles represent the values obtained by solving Equation (12) with an initial-value-problem approach, which is independent of the eigen-mode analysis presented in the text. The density contrast \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\) is chosen to be 100.

Figure 2a indicates that the wave period \(P\) increases monotonically with \(L/R\) in the trapped regime and rapidly settles to some asymptotic value in the leaky one. Likewise, Figure 2b shows that, being identically infinite in the trapped regime, the damping time \(\tau\) also experiences saturation for sufficiently large \(L/R\). In addition, both \(P\) and \(\tau\) increase substantially with increasing \(\mu\) at a given \(L/R\). We note that while the tendency for \(P\) to be larger for steeper density profiles agrees with the study by Nakariakov, Hornsey, and Melnikov (2012), it does not hold in general. Figure 3 in Paper I shows that the opposite occurs for some different profile prescriptions. This means that the largely unknown specific form of the transverse density distribution plays an important role in determining the dispersive properties of sausage modes. Consequently, when the period and damping time of sausage modes are seismologically exploited, the uncertainty in specifying the density profile needs to be considered.

3.2 Applications to Spatially Unresolved QPP Observations

In essence, Figure 2 indicates that the period \(P\) and damping time \(\tau\) of sausage modes can be formally expressed as

$$\begin{aligned} & P_{\mathrm{saus}} =\frac{R}{v_{\mathrm{Ai}}}F_{\mathrm{saus}} \biggl( \frac {L}{R}, \mu, \frac{\rho_{\mathrm{i}}}{\rho_{\mathrm{e}}} \biggr), \end{aligned}$$
(8)
$$\begin{aligned} & \frac{\tau_{\mathrm{saus}}}{P_{\mathrm{saus}}} = G_{\mathrm{saus}} \biggl( \frac{L}{R}, \mu, \frac{\rho_{\mathrm{i}}}{\rho _{\mathrm{e}}} \biggr) . \end{aligned}$$
(9)

We note that the damping-time-to-period ratio \(\tau/P\) is adopted here instead of \(\tau\) itself, the reason being that \(\tau/P\) does not depend on \(R/v_{\mathrm{Ai}}\). Furthermore, the \(L/R\)-dependence disappears for cylinders with large enough \(L/R\).

We first consider the applications of Equations (8) and (9) to spatially unresolved QPP events, for which only \(P\) and/or \(\tau\) can be regarded known. However, the information is missing on both the physical parameters \([v_{\mathrm{Ai}}, \mu, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\) and geometrical parameters \([L, R]\). If a trapped sausage mode is responsible for causing a QPP event, which occurs when the signals do not show clear damping, then only Equation (8) is relevant. This means that any point on a three-dimensional (3D) hypersurface in the 4D space formed by \([R/v_{\mathrm{Ai}}, L/R, \mu, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\) is possible to reproduce the measured \(P\). Even if the signals in a QPP event are temporally decaying, the range of possible parameters that can reproduce the measured \(P\) and \(\tau\) is still too broad to be useful: a 2D surface in the 4D parameter space results. The situation improves if we can assume that the flare loops hosting sausage modes are sufficiently thin such that the \(L/R\)-dependence drops out. Equations (8) and (9) then suggest that for trapped (leaky) modes, we can deduce a 2D surface (1D curve) in the 3D space formed by \([R/v_{\mathrm{Ai}}, \mu , \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\). We note that the idea for deriving 1D inversion curves was first introduced by Arregui et al. (2007) and later explored in e.g., Goossens et al. (2008) and Soler et al. (2014). While resonantly damped kink modes were examined therein, the same idea also applies to leaky sausage modes in thin cylinders, the only difference being that the transverse Alfvén time (\(R/v_{\mathrm{Ai}}\)) replaces the longitudinal one (\(L/v_{\mathrm{Ai}}\)).

Figure 3 presents the 1D curve and its projections (the dashed lines) onto various planes in the \([R/v_{\mathrm{Ai}}, \mu, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\) space, using the QPP event reported in McLean and Sheridan (1973) as an example. For this event, the oscillation period is \(4.3~\mbox{s}\) and the damping-time-to-period ratio is ten. Table 1 presents a set of values read from the solid curve in Figure 3. Of the parameters to be inferred, the transverse Alfvén time \(R/v_{\mathrm{Ai}}\) and density ratio \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\) can be somewhat constrained. To be specific, the pair \([R/v_{\mathrm{Ai}}, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\) reads \([2.94~\mbox{s}, 182]\) when \(\mu=1\), and reads \([1.64~\mbox{s}, 88.2]\) when \(\mu=100\). However, the steepness parameter \(\mu\) is difficult to constrain, since its possible range is too broad. This agrees with Paper I, where we concluded that for spatially unresolved QPPs, the transverse Alfvén time is the best constrained, whereas the steepness (the length of the transition layer in units of loop radius \(l/R\) in that paper) corresponds to the other extreme.

Figure 3
figure 3

Inversion curve (the solid line) and its projections (dashed) in the three-dimensional parameter space spanned by \([\mu, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}, R/v_{\mathrm{Ai}}]\). All points along this curve are equally possible to reproduce the quasi-periodic-pulsation event reported in McLean and Sheridan (1973), where the oscillation period is \(4.3~\mbox{s}\), and the damping-time-to-period ratio is ten.

Table 1 Values of \([\mu, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}, R/v_{\mathrm{Ai}}]\) deduced for the QPP event reported in McLean and Sheridan (1973).

The question then is how to make sense of this seismological inversion. To this end, we may compare our results with what is found with the DR for a step-function density profile (Equation (18) in Paper I). Noting that the \(\mu\)-dependence no longer exists in the step-function case, we find with the measured \(P\) and \(\tau\) that \([R/v_{\mathrm{Ai}}, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}] = [1.62~\mbox{s}, 88.2]\). As expected, this agrees well with what we found for large \(\mu\). However, it differs substantially from the results for small \(\mu\). From this we conclude that, although simple and straightforward, the practice for deducing \([R/v_{\mathrm{Ai}}, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\) using the DR for step-function profiles is subject to substantial uncertainty if the uncertainties in prescribing the transverse density structuring are taken into account. In particular, it may substantially underestimate \(R/v_{\mathrm{Ai}}\). This uncertainty will be carried over to the deduced values of the Alfvén speed and consequently the magnetic field strength, provided that we can further estimate the loop radius \(R\) and internal density \(\rho_{\mathrm{i}}\).

3.3 Applications to Spatially Resolved QPP Observations

We now consider the seismological applications of Equations (8) and (9) to spatially resolved QPP events. In this case, the geometrical parameters \(L\) and \(R\) can be considered known, and only the combination of \([v_{\mathrm{Ai}}, \mu, \rho_{\mathrm{i}}/\rho _{\mathrm{e}}]\) remains to be deduced. It then follows that if a trapped (leaky) mode is presumably the cause of a QPP event, the measured period \(P\) (\(P\) together with the damping time \(\tau\)) allows a 2D surface (1D curve) to be found in the 3D space formed by \([v_{\mathrm{Ai}}, \mu, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\).

Something more definitive can be deduced if a QPP event involves more than just a sausage mode. Similar to Paper I, we examined the case where a fundamental kink mode exists together with a fundamental sausage one, with both experiencing temporal damping. Let \(P_{\mathrm{saus}}\) and \(\tau_{\mathrm{saus}}\) denote the period and damping time of the sausage mode, respectively. Likewise, let \(P_{\mathrm{kink}}\) (\(\tau_{\mathrm{kink}}\)) denote the period (damping time) of the kink mode. Furthermore, we assume that wave leakage leads to the apparent damping of the sausage mode, whereas resonant absorption is responsible for damping the kink mode. We find that \(P_{\mathrm{kink}}\) and \(\tau_{\mathrm{kink}}\) can be formally expressed as

$$\begin{aligned} & P_{\mathrm{kink}} = \frac{L}{v_{\mathrm{Ai}}} F_{\mathrm{kink}} \biggl( \frac {L}{R}, \mu, \frac{\rho_{\mathrm{i}}}{\rho_{\mathrm{e}}} \biggr), \end{aligned}$$
(10)
$$\begin{aligned} & \tau_{\mathrm{kink}} =\frac{L}{v_{\mathrm{Ai}}} H_{\mathrm{kink}} \biggl( \frac {L}{R}, \mu, \frac{\rho_{\mathrm{i}}}{\rho_{\mathrm{e}}} \biggr) . \end{aligned}$$
(11)

To establish the functions \(F_{\mathrm{kink}}\) and \(H_{\mathrm{kink}}\), we adopted the same approach as in Terradas, Oliver, and Ballester (2006). A set of linearized resistive MHD equations (Equations (1) – (5) therein) was solved for the dimensionless complex angular frequency (\(\omega _{\mathrm{kink}} L/v_{\mathrm{Ai}}\)) as an eigenvalue. A uniform resistivity \(\bar{\eta}\) was adopted, resulting in a magnetic Reynolds number \(R_{\mathrm{m}} = v_{\mathrm{Ai}}R/\bar{\eta}\). It turns out that \(\omega_{\mathrm{kink}} L/v_{\mathrm{Ai}}\) does not depend on \(R_{\mathrm{m}}\) when \(R_{\mathrm{m}}\) is sufficiently large, and this saturation value is taken to be the value that \(\omega_{\mathrm{kink}} L/v_{\mathrm{Ai}}\) attains with the input parameters \([L/R, \mu, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\) (see Appendix B for details). We note that \(F_{\mathrm{kink}}\) and \(H_{\mathrm{kink}}\) can also be established with the approach developed by Soler et al. (2013), where a less computationally costly method based on singular series expansion was employed.

With \(P_{\mathrm{saus}}\), \(\tau_{\mathrm{saus}}\), \(P_{\mathrm{kink}}\), and \(\tau _{\mathrm{kink}}\) measured, we find that the number of equations is more than needed, since now there are only three unknowns, \(v_{\mathrm{Ai}}\), \(\mu\), and \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\). In practice, we consider the expression for \(P_{\mathrm{kink}}\) as redundant and use the rest for seismological purposes. As suggested in Paper I, the kink mode period expected from Equation (10) with the deduced parameters can be compared with the measured value. The difference between the two allows us to say, for instance, whether it is safe to identify the oscillating signals with the particular modes. In addition, this difference can also serve as an estimate of the errors of the deduced loop parameters for a given density prescription.

While seemingly fortuitous, QPP events involving multiple modes do occur (e.g., Nakariakov, Melnikov, and Reznikova, 2003; Kupriyanova, Melnikov, and Shibasaki, 2013; Kolotkov et al., 2015). For instance, when analyzing the multiple signals in the QPP event on 14 May 2013, Kolotkov et al. (2015) identified a fundamental fast kink mode with period \(P_{\mathrm{kink}}=100~\mbox{s}\) and damping time \(\tau_{\mathrm{kink}}=250~\mbox{s}\) together with a fundamental sausage mode with \(P_{\mathrm{saus}}=15~\mbox{s}\) and \(\tau_{\mathrm{saus}}=90~\mbox{s}\). In addition, the flare loop hosting the two modes was suggested to be of length \(L=4\times10^{4}~\mbox{km}\) and radius \(R=4\times10^{3}~\mbox{km}\), if the apparent width of the loop is taken as the loop diameter. Now the seismological inversion is rather straightforward and involves only two steps, as illustrated in Figure 4. First, with the aid of Equation (9), we readily derive a curve (the solid curve in Figure 4a) in the \([\mu, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\) plane to be compatible with the measured \(\tau_{\mathrm{saus}}/P_{\mathrm{saus}}\). The internal Alfvén speed \(v_{\mathrm{Ai}}\) for a given pair of \([\mu , \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\) is then found with Equation (8) to agree with the measured \(P_{\mathrm{saus}}\), yielding the dash-dotted curve. Second, we evaluate the kink mode damping time with Equation (11) with a series of combinations \([v_{\mathrm{Ai}}, \mu, \rho_{\mathrm{i}}/\rho _{\mathrm{e}}]\), thereby finding the solid curve in Figure 4b. The intersection of this solid curve with the horizontal dashed line, representing the measured kink mode damping time (\(\tau_{\mathrm{kink}}=250~\mbox{s}\)), then yields that \(\mu= 11.8\), \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}} = 28.8\), and \(v_{\mathrm{Ai}} = 653~\mbox{km}\,\mbox{s}^{-1}\). It is worth stressing that Equation (10) yields an expected kink mode period of \(88~\mbox{s}\) with the measured \(L\) and \(R\) as well as this set of deduced parameters. This is close to what was measured (\(100~\mbox{s}\)), substantiating the interpretation of the long-period signal as the fundamental kink mode, as was done by Kolotkov et al. (2015). Alternatively, this agreement between the two values also suggests that the errors in this inversion procedure are rather moderate.

Figure 4
figure 4

Illustration of the scheme for inverting the two-mode QPP event reported in Kolotkov et al. (2015). The curves in (a) are found by requiring that the damping-time-to-period ratio \(\tau_{\mathrm{saus}}/P_{\mathrm{saus}}\) and period \(P_{\mathrm{saus}}\) for fundamental sausage modes agree with the measured values for a series of given values of \(\mu\). The solid curve in (b) represents the damping time \(\tau_{\mathrm{kink}}\) for fundamental kink modes expected with the values \([\mu, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}, v_{\mathrm{Ai}}]\) given in (a). Its intersection with the horizontal dashed line, representing the measured value for \(\tau_{\mathrm{kink}}\), gives a unique combination of \([\mu, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}, v_{\mathrm{Ai}}]\) labeled by the crosses.

What are the uncertainties of the derived flare loop parameters? Evidently, they come entirely from the uncertainties associated with the unknown specific form of the transverse density structuring. To provide an uncertainty measure, we repeated the afore-mentioned inversion process for all four different density prescriptions in Paper I, where we examined only one profile (the sine profile) and adopted the TTTB approximation to describe \(F_{\mathrm{kink}}\) and \(H_{\mathrm{kink}}\). Now with the pertinent analytical DRs for sausage modes and self-consistent resistive MHD computations for kink modes, we find that the density contrast \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\) is constrained to the range from 28.4 to 31.1, and the internal Alfvén speed \(v_{\mathrm{Ai}}\) lies between 594 and \(658~\mbox{km}\,\mbox{s}^{-1}\). Interestingly, for the \(\mu\)-power profile examined here, the values inferred for \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\) and \(v_{\mathrm{Ai}}\) also lie in these rather narrow ranges. On the other hand, the deduced \(\mu\) value indicates that the density profile across the flare loop in question is rather steep, which also agrees with the ratios of the transition layer width to loop radius (\(0.167 \le l/R \le0.284\)) inferred with the profile prescriptions in Paper I. From this we conclude that at least for the profiles examined in the present study and Paper I, the uncertainties of the inferred profile steepness, density contrast, and internal Alfvén speed are relatively small.

4 Summary

A substantial fraction of quasi-periodic pulsations (QPPs) in the light curves of solar flares is attributed to sausage modes in flare loops. The present study continues the effort we initiated in Chen et al. (2015b, Paper I) to infer flare loop parameters with the measured periods \(P\) and damping times \(\tau\) of fundamental standing sausage modes supported therein. For this purpose we extended the analysis of Paper I to sausage waves in nonuniform, straight, coronal cylinders with arbitrary transverse density profiles comprising a nonuniform inner portion and a uniform external medium. Working in the framework of ideal and cold MHD, we derived an analytical dispersion relation (DR, Equation (3)) and focused on density profiles of a \(\mu\)-power form (Equation (7)). The dispersive properties of fundamental standing modes were examined, together with their potential to infer flare loop parameters.

We found that \(P\) and \(\tau\) in units of the transverse Alfvén time \(R/v_{\mathrm{Ai}}\) depend only on the density contrast \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\), length-to-radius ratio \(L/R\) of coronal cylinders, and the profile steepness \(\mu\). For all profiles examined in both this study and Paper I, when the rest of the parameters are fixed, \(P\) (\(\tau\)) in units of \(R/v_{\mathrm{Ai}}\) increases (decreases) with increasing \(L/R\) and tends to some saturation value when \(L/R\) is sufficiently large. For spatially unresolved QPPs, we showed that a curve in the 3D space formed by \(R/v_{\mathrm{Ai}}\), \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\), and \(\mu\) can almost be deduced. This occurs when we can assume that \(L/R \gg1\) beforehand. Applying this inversion procedure to the event reported by McLean and Sheridan (1973), we found that \(R/v_{\mathrm{Ai}}\) is the best constrained, whereas the steepness parameter is the least constrained. For spatially resolved QPPs, we showed that while geometric parameters of flare loops are available, the inversion problem remains under-determined. However, when an additional mode co-exists with the fundamental sausage mode, the full information on the unknowns, \([v_{\mathrm{Ai}}, \mu, \rho_{\mathrm{i}}/\rho_{\mathrm{e}}]\), can be inferred. In this case, the inversion problem may become over-determined. Applying this idea to a recent QPP event where temporally decaying kink and sausage modes were identified, we found that \(v_{\mathrm{Ai}}\), \(\rho_{\mathrm{i}}/\rho_{\mathrm{e}}\), and the profile steepness can be constrained to rather narrow ranges.

The discussions on the limitations to our inversion procedures as presented in Paper I also apply here and are not repeated. Instead, we stress the great potential of using multi-mode QPP measurements to determine flare loop parameters rather precisely, the internal Alfvén speed in particular. To this end, not only modes of distinct nature (e.g., a fundamental kink mode co-existing with a sausage one) are useful, modes of the same nature but with different longitudinal node numbers work as well. While fundamental kink modes and their harmonics have been seismologically exploited (see e.g., the review by Andries et al., 2009), serious studies using sausage modes need to be conducted.

Before closing, we note that Bayesian techniques have been successfully applied to the inference of density structuring transverse to coronal loops hosting resonantly damping kink modes (Asensio Ramos and Arregui, 2013; Arregui, Asensio Ramos, and Pascoe, 2013; Arregui, Soler, and Asensio Ramos, 2015). With such techniques, the errors in the measurements of kink mode periods and damping times can be properly propagated, and the plausibility of a density profile prescription can be assessed. When no particular prescription is favored, approaches like model-averaging can be employed to yield an evidence-averaged inference. While so far the applications of such techniques have been primarily focused on kink modes, similar ideas are expected to be equally applicable to sausage modes. For this purpose, the DRs derived here and in Paper I should be useful.