Keywords

These keywords were added by machine and not by the authors. This process is experimental and the keywords may be updated as the learning algorithm improves.

1 What This Paper Is About and What It Is Not

Viscoelasticity is of current interest to geology. Typically geologists investigate the temporal evolution of deformation within the Earth’s outer crust caused by earthquakes or other gravitational load shifts such as melting ice, using viscoelastic material models, e.g., Campbell (1974), Ragazzo and Ruiz (2015) and Tanaka et al. (2009).

The present paper is not going in this direction at all. Rather it is a idealistic continuum approach toward an understanding of the genesis of terrestrial planets and the subsequent state of deformation in a large self-gravitating object. However, there is also a certain esthetic aspect in the solution we are about to present, and in order to quote Keats we may say that “beauty is truth, truth beauty.” More specifically, our result is a follow-up on the classical solutions found by the great A.E.H. Love for a self-gravitating linear-elastic sphere, see Love (1892, 1906, 1927). We shall extend his beautiful formulae to a linear-elastic model of the Kelvin–Voigt type. In other words, we will explore the temporal development of the static linear-elasticity solution of a self-gravitating terrestrial planet. In particular, we shall look at the temporal evolution of the Love radius, i.e., the position of the transition zone between compression and tension within a self-gravitating “solid” sphere. This may even be of practical use, since it is related to damage during the early stages of a developing terrestrial planet. However, we shall not endeavor to investigate this in full quantitative detail, at least not here. Surprisingly, our results will be of closed form, thanks to the efforts of one of the authors in a completely different field of research, cf. Frelova (2016). This shows the power of continuum theory: Everything is connected, a maxim we chose to start our salute to our esteemed colleague Holm Altenbach!

2 Literature Review and Putting the Problem into Perspective

Today it is a commonly accepted opinion that terrestrial planets, such as Mercury, Venus, Earth, and Mars, but also other huge solid celestial objects, specifically the Moon,Footnote 1 are the result of a coagulation process of “rocky” matter, a.k.a. “planetesimals,” to form so-called “protoplanets” during the early stages of the developing solar system, cf., Wetherill (1990). In order to quote Lissauer (1993), pg. 134: “... in this picture, planet formation is fundamentally different from star formation in that planetary growth begins with the accumulation of solid bodies, with the accretion of substantial amounts of gas occurring after a planet becomes sufficiently massive ...” and, pg. 136, “... These planetesimals continue to agglomerate via pairwise mergers. ... Growth via binary collisions proceeds until the protoplanets become dynamically isolated from each other.”

Hearing all this, we might conclude that the mathematical modeling of the genesis of a planet is exclusively numerical and within the field of discrete mechanics or (better) discrete systems, since there will be thermodynamics aspects involved, see, e.g., Kenyon (2006). However, it is always wise to look at a problem from different angles and, consequently, we promote the continuum perspective in what follows.

Let us consider the following scenario: A spherical, initially homogeneous, unstressed sphere (the planet in statu nascendi) undergoes self-gravity. We must ask as to whether static equilibrium is possible and how it is reached? Two rather idealized scenarios come to mind.

First, imagine that gravity is “suddenly switched on.” Then, we will essentially face a situation similar to that of a moving masspoint connected to a linear-elastic Hookean spring: Due to the inertial terms in the equation of motion and due to the potential of a linear oscillator pertinent to a radially symmetric, self-gravitating sphere, this sphere will begin to shrink below the radius determined by static equilibrium of forces. While doing so, stress-related forces will build up so that the sphere will finally start to rebounce. Provided there is no dissipation it will reach its initial radius again. This will happen over and over if we assume the material of the sphere to be perfectly elastic without internal friction and without heat conduction, so that isothermal conditions prevail. In other words, without dissipation there will be a constant exchange between the elastic energy, the gravitational potential, and the kinetic energies: The motion of the self-gravitating matter would never come to a standstill. Of course, in the real world there are dissipative processes acting. The shrinking will be accompanied by dissipation in terms of viscoelastic or viscoplastic deformation, and there will be heat conduction. All of this will, in the end, bring motion to a standstill, and the sphere will arrange itself in thermomechanical equilibrium, i.e., there will be equilibrium of gravitational and inner, stress-related forces in a state of homogeneous temperature. It should be mentioned that the final equilibrium state of a self-gravitating sphere has been modeled in closed mathematical form at the end of the nineteenth century by Love (1892), who used linear elasticity at small deformations for this purpose. The interested reader will find detailed information about Love’s solution and interpretation in the Appendix.

In conclusion, we shall not attempt to model the dynamic transition toward that equilibrium for various reasons. First of all, its treatment would be fully numerical based on large deformations expressed in terms of velocities. This makes it difficult if not impossible to compare it to Love’s analysis of equilibrium, which was based on small strains. Moreover, choosing an adequate numerical technique would be required. Surely there will be more than one, all of them with certain pros and cons. Finally, the question which initial conditions are appropriate is difficult to answer, since gravity is not simply “switched on” but always present. Hence in terms of capturing reality our dynamic continuum model could not seriously compete with the discrete mechanics approach of planetesimal masses bouncing into each other, sticking together, and finally forming a primordial planet that relaxes stress- and displacement-wise under the influence of their mutual gravitational attraction. In short, the fully dynamic continuum model requires too much effort for little gain.

For all these reasons, we shall eat humble pie and turn alternatively to a quasistatic treatment instead. This way inertial forces in the equations of motion can be neglected and the self-gravitating sphere will quasistatically and isothermally move into its final state of deformation. Such a situation is frequently conjured up in so-called \(p\,\mathrm {d}V\)-thermodynamics, for example, if we allow the pressure on a piston to change very slowly so that the gas which is trapped in the corresponding container has time to accommodate pressure- and temperature-wise. However, in our approach the time parameter will enter through a viscoelastic model used to connect stresses, strains, and their corresponding rates. More specifically, in order to be able to study the temporal development of the solution for the displacements, strains, and stresses toward Love’s closed-form solutions we will make use of a linear viscoelastic model of the Kelvin–Voigt type, i.e., small deformation theory will reappear.

Fig. 1
figure 1

Stress- versus displacement-controlled viscoelastic experiments (see text)

In this context recall the two fundamental types of quasistatic experiments always mentioned in combination with quasistatic, linear viscoelasticity (see Fig. 1)Footnote 2: In the first one a linear-viscoelastic strip is suddenly subjected to a constant “dead load,” i.e., a constant uniaxial tensile stress (the “cause”), \(\sigma _0\), is prescribed. Under such circumstances we also speak of load-control. The “effect” consists of an elastic strain, \(\epsilon _\mathrm {i}\), instantaneously built up. After that the strip gradually creeps quasistatically toward its final total strain, \(\epsilon _\mathrm {f}\). The counterpart to this experiment consists of prescribing a strain of a fixed amount (the “cause”), \(\epsilon _0\), and to observe the stress response (the “effect”). This is what we call a displacement-controlled test. It turns out that the stress response immediately overshoots to a high level, \(\sigma _\mathrm {i}\), and is then reduced by creeping quasistatically toward a final lower value, \(\sigma _{\mathrm {f}}\). This time we speak of stress-relaxation.

In the following section we shall state and solve the linear-viscoelastic problem for a self-gravitating sphere mathematically and study the behavior of the corresponding solution which, surprisingly, will also be of closed-form. Moreover, we shall also investigate as to whether this fits into the traditional pattern of stress or strain controlled experiments.

3 A Viscoelastic Model of Self-gravitation

3.1 Viscoelasticity of the Kelvin–Voigt Type

Recall the 1D representation of the so-called Kelvin–Voigt model: A Hookean spring and a dashpot are arranged in parallel: Fig. 2. If we apply a displacement, \(\delta \), at the outer points of this rheological model it will be transferred equally to the spring and to the dashpot, \(\delta =\delta _1=\delta _2\), whereas the resulting force is the sum of the forces due to both elements, \(F=F_1+F_2\). In strength-of-materials-terminology we may say that the strains and, hence, the strain rates are equal, \(\epsilon =\epsilon _1=\epsilon _2\Rightarrow \dot{\epsilon }=\dot{\epsilon }_1=\dot{\epsilon }_2\), whereas the stresses are additive, \(\sigma =\sigma _1+\sigma _2\), and where the dot refers to a time derivative. The spring is now modeled by Hooke’s law, \(\sigma _1=E\epsilon _1\), and the dashpot by a Newton–Navier–Stokes relationship, \(\sigma _2=\eta \,\dot{\epsilon }_2\). If we combine these equations, we arrive at:

$$\begin{aligned} \begin{gathered} \sigma =E\left( \frac{\eta }{E}\dot{\epsilon }+\epsilon \right) . \end{gathered}\end{aligned}$$
(1)

We can then introduce a strain-based relaxation time, , which will come in handy once we turn to dimensionless equations.

Fig. 2
figure 2

The Kelvin–Voigt model of 1D linear viscoelasticity

One-dimensional rheological models allow us to find three-dimensional analogues if we accept the daring idea of replacing the 1D stress and strain simply by stress and strain tensors. We proceed to illustrate this idea for the case of the Kelvin–Voigt model, Eq. (1), and write:

$$\begin{aligned} \begin{gathered} \varvec{\sigma }=\varvec{\sigma }_1+\varvec{\sigma }_2\,,\quad \varvec{\epsilon }=\varvec{\epsilon }_1=\varvec{\epsilon }_2. \end{gathered}\end{aligned}$$
(2)

Now we assume that the planet can be modeled as a linear, isotropic medium. We therefore consider the following customary constitutive equations for linear elasticity and viscosity during further analysis (the acronyms “dil” and “dev” refer to dilatoric and deviatoric parts of the strain (rate) tensors, respectively):

$$\begin{aligned} \begin{gathered} \varvec{\sigma }_{1}=3k\,\varvec{\epsilon }_1^{\mathrm {dil}}+2\mu \,\varvec{\epsilon }_1^{\mathrm {dev}}\,,\quad \varvec{\sigma }_{2}=3\eta '\,\dot{\varvec{\epsilon }_2}^{\mathrm {dil}}+2\eta \,\dot{\varvec{\epsilon }_2}^{\mathrm {dev}}\,, \end{gathered}\end{aligned}$$
(3)

where k and \(\mu \) refer to the bulk modulus and the shear modulus, respectively. Moreover, \(\eta '\) and \(\eta \) are known as coefficients of bulk and shear viscosity.

If we now combine Eqs. (2) and (3) suitably we finally arrive at the following relation:

$$\begin{aligned} \begin{gathered} \varvec{\sigma }=3k\left( \frac{\eta '}{k}\dot{\varvec{\epsilon }}^{\mathrm {dil}}+\varvec{\epsilon }^{\mathrm {dil}}\right) +2\mu \left( \frac{\eta }{\mu }\,\dot{\varvec{\epsilon }}^{\mathrm {dev}}+\varvec{\epsilon }^{\mathrm {dev}}\right) . \end{gathered}\end{aligned}$$
(4)

Hence, in principle, we must distinguish between two different relaxation times:

$$\begin{aligned} \begin{gathered} \tau _{\mathrm {v},\epsilon }=\frac{\eta '}{k},\quad \tau _{\mathrm {s},\epsilon }=\frac{\eta }{\mu }\,, \end{gathered}\end{aligned}$$
(5)

where the indices v and s are supposed to remind us of the dilatoric (\(\varvec{\mathrm {v}}\)olumetric) and deviatoric (\(\varvec{\mathrm {s}}\)hear) parts, and the index \(\epsilon \) of the strain-related relaxation process.

However, it is known that the bulk viscosity is a rather elusive parameter and very difficult to measure, see Gad-el Hak and Bandyopadhyay (1995). Therefore, we will neglect it in what follows and obtain from the previous equations because of \(\varvec{\epsilon }^\mathrm {dil}:=\tfrac{1}{3}{{\mathrm{{\textsf {Tr}}}}}\varvec{\epsilon }\varvec{\mathrm {I}}\) and \(\varvec{\epsilon }^\mathrm {dev}:=\varvec{\epsilon }-\varvec{\epsilon }^\mathrm {dil}\):

$$\begin{aligned} \begin{gathered} \varvec{\sigma } = 3k\,\varvec{\epsilon }^\mathrm {dil}+2\mu \,\varvec{\epsilon }^\mathrm {dev}+2\eta \,\dot{\varvec{\epsilon }}^\mathrm {dev}\equiv \lambda \,{{\mathrm{{\textsf {Tr}}}}}\varvec{\epsilon }\,{\text {I}}+2\mu \,\varvec{\epsilon }+2\eta \,\left( \dot{\varvec{\epsilon }}-\frac{1}{3}{{\mathrm{{\textsf {Tr}}}}}\dot{\varvec{\epsilon }}\,\varvec{\mathrm {I}}\right) , \end{gathered}\end{aligned}$$
(6)

\(\varvec{\mathrm {I}}\) being the unit tensor.

Our main objective is to determine the displacement, \(\varvec{u}\), in spherical coordinates (for obvious reasons). We assume perfect spherical symmetry, hence \(\varvec{u}=u_r(r)\varvec{e}_r\), \(u_r\) being its radial component and \(\varvec{e}_r\) being the radial unit vector. All necessary equations will be written in spherical coordinates. Specifically, we recall Eqs. (37) and (38) from the Appendix, which we complement by:

$$\begin{aligned} \begin{gathered} \dot{\epsilon }_{rr}=\dot{u}'_{r}\; ,\qquad \dot{\epsilon }_{\vartheta \vartheta }\equiv \dot{\epsilon }_{\varphi \varphi }=\frac{\dot{u}_{r}}{r}\; ,\qquad \dot{\epsilon }_{r\vartheta }=\dot{\epsilon }_{r\varphi }=\dot{\epsilon }_{\vartheta \varphi }\equiv 0\;. \end{gathered}\end{aligned}$$
(7)

Then we obtain analogously to Eq. (39):

$$\begin{aligned} \sigma _{rr}&=(\lambda +2\mu )u'_{r}+2\lambda \frac{u_{r}}{r}+\frac{4}{3}\eta \left( \dot{u}_r'-\frac{\dot{u}_r}{r}\right) , \nonumber \\ \sigma _{\vartheta \vartheta }&\equiv \sigma _{\varphi \varphi }=\lambda u'_{r}+2(\lambda +\mu )\frac{u_{r}}{r}+\frac{2}{3}\eta \left( \frac{\dot{u}_r}{r}-\dot{u}'_r\right) ,\; \\ \sigma _{r\vartheta }&=\sigma _{r\varphi }=\sigma _{\vartheta \varphi }\equiv 0. \nonumber \end{aligned}$$
(8)

The equilibrium conditions (34) hold and we arrive similarly to Eq. (40) at the following Partial Differential Equation (PDE) for \(u_r(r,t)\):

$$\begin{aligned} \begin{gathered} u''_{r}+2\frac{u'_{r}}{r}-2\frac{u_{r}}{r^2}+\frac{4}{3}\frac{\eta }{\lambda +2\mu }\left( \dot{u}''_{r}+2\frac{\dot{u}'_{r}}{r}-2\frac{\dot{u}_{r}}{r^2}\right) =\frac{4\pi \rho ^2_0 G}{3(\lambda +2\mu )}r, \end{gathered}\end{aligned}$$
(9)

where the dot means differentiation w.r.t time, t, and the dash differentiation w.r.t. position, r.

This PDE must be solved in combination with two boundary conditions and one initial condition. We will study the case of a viscoelastic sphere of outer radius, \(r_\mathrm {o}\). The boundary conditions state that the displacement stays finite and vanishes in the center and that there is no traction at the outer boundary, \(r_\mathrm {o}\):

$$\begin{aligned} \begin{gathered} u_{r}(r=0,t)=0\;,\quad \sigma _{rr}(r=r_\mathrm {o},t)=0. \end{gathered}\end{aligned}$$
(10)

3.2 Solution in Dimensionless Form

Analogously to Eqs. (47) and (49)\(_1\) we define

$$\begin{aligned} \begin{gathered} u\equiv u(x,\tau )=\frac{u_r}{r_\mathrm {o}}\;,\quad x=\frac{r}{r_\mathrm {o}}\;,\quad \tau =\frac{\lambda +2\mu }{\eta }t\;,\quad \alpha =\frac{8\pi G \rho ^2_0 r^2_\mathrm {o}}{3(\lambda +2\mu )}. \end{gathered}\end{aligned}$$
(11)

Then the PDE (9) assumes the form:

$$\begin{aligned} \begin{gathered} u''+2\frac{u'}{x}-2\frac{u}{x^2}+\frac{4}{3}\left( \dot{u}''+2\frac{\dot{u}'}{x}-2\frac{\dot{u}}{x^2}\right) =\frac{\alpha }{2}x, \end{gathered}\end{aligned}$$
(12)

where the dot now refers to differentiation w.r.t. dimensionless time, \(\tau \), and the dash means differentiation w.r.t. dimensionless position, x.

The nonvanishing stresses are normalized by \(\lambda +2\mu \) (identified by a tilde) and read:

$$\begin{aligned} \tilde{\sigma }_{rr}&=u'+\frac{2\nu }{1-\nu }\frac{u}{x}+\frac{4}{3}\left( \dot{u}'-\frac{\dot{u}}{x}\right) , \\ \tilde{\sigma }_{\vartheta \vartheta }&\equiv \tilde{\sigma }_{\varphi \varphi }=\frac{\nu }{1-\nu } u'+\frac{1}{1-\nu }\frac{u}{x}-\frac{2}{3}\left( \dot{u}'-\frac{\dot{u}}{x}\right) . \nonumber \end{aligned}$$
(13)

The boundary conditions (10) take the following form:

$$\begin{aligned} \begin{gathered} u(0,\tau )=0\;, \end{gathered}\end{aligned}$$
(14)
$$ \tilde{\sigma }_{rr}(1,\tau )\equiv u'(1,\tau )+\frac{2\nu }{1-\nu }u(1,\tau )+\frac{4}{3}\bigg [\dot{u}'(1,\tau )-\dot{u}(1,\tau )\bigg ]=0,\;\; $$

and the initial condition reads:

$$\begin{aligned} \begin{gathered} u(x,0)=0\;,\quad x\in [0, 1], \end{gathered}\end{aligned}$$
(15)

this is to say that we expect no displacements initially, because “gravitation has just been switched on at \(\tau =0\).”

We solve the PDE (12) by mapping it onto Laplace space w.r.t. time \(\tau \leftrightarrow s\) and then finding a solution of the corresponding Ordinary Differential Equation (ODE). The Laplace transform of the displacement will be identified by a bar, \(\bar{u}=\bar{u}(x,s)\), and we may write according to the usual rules of Laplace transforms:

$$\begin{aligned}&\left( 1+\frac{4}{3}s\right) \left( \bar{u}''(x,s)+2\frac{\bar{u}'(x,s)}{x}-2\frac{\bar{u}(x,s)}{x^2}\right) -\nonumber \\&\quad \frac{4}{3}\left[ u''(x,0)+2\frac{u'(x,0)}{x}-2\frac{u(x,0)}{x^2}\right] =\frac{\alpha x}{2}\frac{1}{s}. \end{aligned}$$
(16)

The term in brackets drops out. We can give two reasons for that. First, there is the initial condition (15), according to which the displacement (and all its derivatives) shall vanish initially. Second, we note that this very term represents the (stationary) ODE of the gravitational problem provided gravitation is not present, see Eq. (40), which is zero to begin with. The solution of the remaining ODE for \(\bar{u}(x,s)\) is completely analogous to the one presented in Eq. (41). We may write:

$$\begin{aligned} \begin{gathered} \bar{u}(x,s)=Ax+\frac{B}{x^2}+\frac{\alpha }{20}x^3\frac{1}{s\left( 1+\tfrac{4}{3}s\right) } , \quad 0\le x\le 1, \end{gathered}\end{aligned}$$
(17)

In order to determine the constants of integration we have to transform the boundary conditions (14) into Laplace space as follows:

$$\begin{aligned} \begin{gathered} \bar{u}(0,s)=0\;, \end{gathered}\end{aligned}$$
(18)
$$ \left( 1+\frac{4}{3}s\right) \bar{u}'(1,s)+\left( \frac{2\nu }{1-\nu }-\frac{4}{3}s\right) \bar{u}(1,s)-\frac{4}{3}\bigg [u'(1,0)- u(1,0)\bigg ]=0. $$

For the same reasons as before the term in parentheses in the second equation drops out. The first equation requires us to put \(B=0\). Moreover, the remaining linear equation for A in Eq. (17)\(_2\) can be solved and, after back transform into real time space, the final result reads as follows:

$$\begin{aligned} \begin{gathered} u(x,\tau =0)=0\;, \end{gathered}\end{aligned}$$
(19)
$$ u(x,\tau >0)=-\frac{\alpha }{20}x\left[ \frac{3-\nu }{1+\nu }-x^2\right] \left[ 1-\mathrm {exp}\left( -\tfrac{3}{4}\tau \right) \right] -\frac{\alpha }{10}\frac{1-\nu }{1+\nu }x\,\mathrm {exp}\left( -\tfrac{3}{4}\tau \right) . $$

Note that special attention has been given to the case \(\tau =0\): If we consider the limit case \(\tau \rightarrow 0\) we find a nonvanishing initial displacement. Moreover, it can be seen that the initial and boundary conditions from Eqs. (14), (15) are indeed satisfied. For \(\tau \rightarrow \infty \) the stationary relation shown in Eq. (48) is obtained.

We are now in a position to determine the dimensionless stresses from Eq. (13):

$$\begin{aligned} \tilde{\sigma }_{rr}&=-\frac{\alpha }{20}\left( 1-x^2\right) \left[ \frac{3-\nu }{1+\nu }-\frac{1+\nu }{1-\nu }\mathrm {exp}\left( -\frac{3}{4}\tau \right) \right] , \\ \tilde{\sigma }_{\vartheta \vartheta }&\equiv \tilde{\sigma }_{\varphi \varphi }=-\frac{\alpha }{20}\frac{3-\nu }{1-\nu }\left[ 1-\frac{1+3\nu }{3-\nu }x^2-\frac{1+\nu }{3-\nu }\left( 1-2x^2\right) \mathrm {exp}\left( -\frac{3}{4}\tau \right) \right] . \nonumber \end{aligned}$$
(20)

It is easy to see that in the limit \(\tau \rightarrow \infty \) the stresses of the stationary solution from Eq. (52) result.

3.3 Evaluation and Discussion of the Results

Figure 3 depicts the temporal evolution of the displacement as a function of radial distance in dimensionless form as predicted by Eq. (19) for the choice \(\nu =0.3\). Note that immediately after “gravity has been switched on” the dependence is nearly linear.Footnote 3 Consequently, the minimum is located at the outer radius \(x=1\). It is an edge minimum and not a “true” minimum with vanishing derivative.

Fig. 3
figure 3

Temporal development of the displacement as a function of radial position (see text)

In this context recall the notion of the “Love radius.” It indicates the position where the radial strains within a self-gravitating sphere changes sign and it was first discovered by A.E.H. Love. In equilibrium this (normalized) position is given by:

$$\begin{aligned} \begin{gathered} x_\mathrm{{Love}}=\sqrt{\frac{3-\nu }{3(1+\nu )}}, \end{gathered}\end{aligned}$$
(21)

and the details of the derivation of the formula can be found in the Appendix.

Also recall that in the present case the radial strain is nothing else but the derivative of the radial displacement w.r.t. position, i.e., the slope to that curve. Moreover, the Love radius is defined by a true minimum of the radial displacement with zero slope. Consequently, in the transient case, a tensile region does not exist initially. It takes a certain while until the prominent feature of a true minimum corresponding to the location of the Love radius evolves.

Fig. 4
figure 4

Temporal development of the Love radius (see text)

We can obtain the location of the Love radius by (formal) differentiation of the displacement shown in Eq. (19) w.r.t. x. The result is:

$$\begin{aligned} \begin{gathered} x_{\mathrm {Love}}=\frac{1}{\sqrt{3}}\sqrt{\frac{3-\nu }{1+\nu }+\frac{1-\nu }{1+\nu }\frac{1}{\mathrm {exp}\left( \tfrac{3}{4}\tau \right) -1}}. \end{gathered}\end{aligned}$$
(22)

It is easily seen that this expression tends to the “elastic” Love radius shown in Eq. (21) if \(\tau \) goes to infinity (Fig. 4).

Finally, Fig. 5 presents the (dimensionless) stresses. It is noteworthy that the radial stress component relaxes monotonically without a qualitative change in the shape of the curve. This is not so for the angular stresses, whose minima switch from \(x=1\) to \(x=0\) as time goes on.

Based on the results presented in Figs. 3 and 5, we must conclude that the process of relaxation in a self-gravitating terrestrial planet is not as simple as in the textbook example of a load and displacement-controlled viscoelastic strip shown in Fig. 1. This is due to the fact that we face a three-dimensional state of stress after “switching on” a spatially varying body force.

Fig. 5
figure 5

Temporal development of the stresses (see text)

4 Outlook and Conclusions

The main objective of this paper was to present an analysis of the temporal development of the displacements, strains, and stresses in a self-gravitating sphere. The model was based on a radially symmetric linear viscoelastic constitutive model of the Kelvin–Voigt type. An analytical solution was found based on Laplace transforms. It was shown how the displacement and stresses relax to the stationary linear-elastic solution, originally due to Love, which was also reviewed in an appendix. In particular it was shown that the so-called Love radius, which marks the transition between the regions of compressive and tensile strain, does not exist in the early stages. It takes some time to develop.

In future work we will investigate alternative viscoelastic models, for example a generalization of the Zener type. We will also attempt to predict the relaxation time scales based on recent measurements of the viscosity of (liquid) iron and igneous rock.