1 Introduction

Let X be a separable Hilbert space and \(A:D(A)\subset X\rightarrow X\) be an unbounded linear, closed, densely defined, self-adjoint and for some \(a>0\), satisfies \(\langle Au,u\rangle \geqslant a \langle u,u\rangle \) or all \(u\in D(A)\). We also assume that A has compact resolvent on X. It is well-known that with those hypotheses, we can define the fractional power \(A^{\alpha }\) of order \(0<\alpha <1\) according to [2] and [11], as a closed linear operator, see e.g. [10, 11, 13, 18]. Denote by \(X^{\alpha }=D(A^{\alpha })\) for \(0\leqslant \alpha \leqslant 1\) (taking \(A^0:=I\) on \(X^0:=X\) when \(\alpha =0\)). Recall that \(X^{\alpha }\) is dense in X for all \(0\leqslant \alpha \leqslant 1\), for details see [2, Theorem 4.6.5]. The fractional power space \(X^\alpha \) endowed with the norm \(\Vert \cdot \Vert _{X^\alpha }:=\Vert A^{\alpha } \cdot \Vert _X\) is a Banach space. With this notation, we have \(X^{-\alpha }=(X^\alpha )'\) for all \(\alpha >0\), see [2, 26, 27] for the characterization of the negative scale.

In this paper we consider the fourth-order linear equation in time

$$\begin{aligned} u_{tttt}+A^{\frac{2}{5}}u+aA^{\frac{1}{5}}u_{ttt}+bA^{\frac{4}{5}}u_{tt}+cAu_t=0, \end{aligned}$$
(1.1)

where \(a,b,c\geqslant 0\).

We discuss here the well-posedness of Eq. (1.1) in terms of the scalar parameters ab and c considering the theory of strongly continuous bounded linear operators on suitable phase spaces.

In recent years higher-order evolution equations in time have attracted the attention of many researchers. This can be explained by the sensitivity of the conditions of well-posedness and regularity of solutions of these equations in infinite-dimensional Banach spaces, see e.g. [3,4,5, 8, 9, 22], see also the Moore-Gibson-Thompson (MGT) equations, see e.g. [1, 6, 7, 12, 15,16,17, 19,20,21, 24, 25] and reference therein.

We also observe that it is possible to find in the literature some works dedicated to fourth-order equations in time motivated by MGT equations, see e.g. [1, 7, 16, 19,20,21]; in these papers are made formulations for fourth-order equations of the MGT type, and results of blow-up of solutions, well-posedness, stability and regularity of solutions are obtained.

Our best knowledge indicates that no information is known for models of the type (1.1) concerning the well-posedness and regularity of solutions. This paper makes this point. Furthermore, abstract differential equations of order greater than two are generally ill-posed in the sense of the theory of semigroups of bounded linear operators. However, we identify cases where (1.1) is well-posed. Namely, cases in which this equation is associated with a strongly continuous semigroup and instances in which it is associated with an analytic semigroup.

We will divide our problem into two cases: \(c>0\) and \(c=0\). First of all, note that our linear problem (1.1) can be rewritten as a first-order abstract system of the form

$$\begin{aligned} U_t+S_{a,b,c}U=0,\quad t>0, \end{aligned}$$

on the phase space

$$\begin{aligned} Y=X^{1}\times X^{\frac{4}{5}}\times X^{\frac{3}{5}}\times X, \end{aligned}$$

endowed with the usual inner product, where \(U=\left[ {\begin{matrix} u\\ u_t\\ u_{tt}\\ u_{ttt} \end{matrix}}\right] \), and the matrix operator \(S_{a,b,c}\) can be seen as a unbounded linear operator defined by

$$\begin{aligned} D(S_{a,b,c})=X^1 \times X^1 \times X^{\frac{4}{5}} \times X^{\frac{1}{5}} \end{aligned}$$

and

$$\begin{aligned} S_{a,b,c}=\begin{bmatrix} 0 &{}\quad -I &{}\quad 0 &{}\quad 0\\ 0 &{}\quad 0 &{}\quad -I&{}\quad 0 \\ 0&{}\quad 0&{}\quad 0&{}\quad -I\\ A^{\frac{2}{5}}&{}\quad cA &{}\quad bA^{\frac{4}{5}}&{}\quad aA^{\frac{1}{5}} \end{bmatrix} \end{aligned}$$

Namely, the unbounded linear operator \(S_{a,b,c}:D(S_{a,b,c})\subset Y\rightarrow Y\) is closed and densely defined.

It is easily seen that \(\lambda \in \rho (-S_{a,b,c})\) if and only if \(\lambda ^4I+a\lambda ^3 A^{\frac{1}{5}}+b\lambda ^2 A^{\frac{4}{5}}+c\lambda A+A^{\frac{2}{5}}\) is bijective. Consequently, if \(\lambda \in \mathbb {C}\) is such that

$$\begin{aligned} \lambda ^4+a\mu ^{\frac{1}{5}}\lambda ^3+b\mu ^{\frac{4}{5}}\lambda ^2+c\mu \lambda +\mu ^{\frac{2}{5}}=0 \end{aligned}$$

for some \(\mu \in \sigma (A)\), then \(\lambda \in \sigma (-S_{a,b,c})\). Since \(\sigma (A)\) is countably infinite, we will consider the characteristic polynomials”

$$\begin{aligned} p_n(\lambda )=\lambda ^4+a\mu _n^{\frac{1}{5}}\lambda ^3+b\mu _n^{\frac{4}{5}}\lambda ^2+c\mu _n\lambda +\mu _n^{\frac{2}{5}}. \end{aligned}$$
(1.2)

with \(\mu _n\in \sigma (A)\). Recall that \(\mu _n>0\) and \(\mu _n\rightarrow \infty \) as \(n\rightarrow \infty \).

Note that the discriminant of (1.2) is given by

$$\begin{aligned} {\begin{matrix} \Delta _n&{}=(-4b^3c^2)\mu _n^{\frac{22}{5}}+(-27c^4+18abc^3+a^2b^2c^2)\mu _n^4+(16b^4-4a^3c^3 )\mu _n^{\frac{18}{5}}+(144a^2b)\mu _n^{\frac{17}{5}}\\ &{}+(144bc^2-80ab^2c-4a^2b^3)\mu _n^{\frac{16}{5}}+(18a^3bc-6a^2c^2)\mu _n^{\frac{14}{5}}-(128b^2)\mu _n^{\frac{12}{5}}-(192ac)\mu _n^2\\ &{}-27a^4\mu _n^{\frac{8}{5}}+256\mu _n^{\frac{6}{5}}, \end{matrix}} \end{aligned}$$

see [14, Example on the pages 192-193].

We can summarize the behavior of \(\Delta _n\) when \(n>>0\) as follows (Table 1).

See Tables 1, 2, 3 below

Table 1 Signal of the discriminant \(\Delta _n\)

This is a key result in our analysis since a geometric localization of the points of the spectral set of \(-S_{a,b,c}\) can provide us with information about the generation of strongly continuous semigroups of bounded linear operators associated with the problem (1.1), see e.g. [23, Chapter 1, Corollary 3.8].

This work is organized as follows. In Sect. 2 we consider the case \(c>0\) and we treat three subcases \(c=ab\), \(c>ab\), and \(c<ab\). Finally, in Sect. 3 we consider the case \(c=0\) and we treat three subcases \(a=b=c=0\), \(a\ne 0\) and \(b=c=0\), \(b\ne 0\) and \(a=c=0\), and \(a,b\ne 0\) and \(c=0\).

The tables below summarize our results, where \(\checkmark \) means that it is possible to generate the respective strongly continuous semigroups of bounded linear operators on Y, and \(\times \) means that it is not possible to generate the respective strongly continuous semigroups of bounded linear operators on Y.

For \(c>0\), we have the following results (Table 2):

Table 2 Case \(c>0\)

For \(c=0\), we have the following results (Table 3):

Table 3 Case \(c=0\)

2 Case \(c>0\)

Since \(\Delta _n<0\) for large n when \(c>0\) we guarantee the existence of two complex conjugated roots \(z_n\) and \(\overline{z_n}\) and two real roots \(x_n\) and \(y_n\). The following lemma shows the behavior of one of the real roots for the polynomial \(p_n\) in (1.2).

Lemma 1.1

If \(c>0\), then \(y_n\rightarrow 0\) as \(n\rightarrow +\infty \).

Proof

Given \(\epsilon >0\) we have

$$\begin{aligned} p_n(-\epsilon )=\epsilon ^4-a\mu _n^{\frac{1}{5}}\epsilon ^3+b\mu _n^{\frac{4}{5}}\epsilon ^2-c\mu _n\epsilon +\mu _n^{\frac{2}{5}}. \end{aligned}$$

It is easy to see that \(p_n(-\epsilon )\rightarrow -\infty \) as \(n\rightarrow +\infty \). Since \(p_n(0)>0\) for any n, we guarantee that \(y_n\in (-\epsilon ,0)\) for large n; that is, \(y_n\rightarrow 0\) as \(n\rightarrow +\infty \).

Corollary 1.2

If \((x_n+a\mu _n^{\frac{1}{5}})\rightarrow K\) as \(n\rightarrow +\infty \), then \(Re(z_n)\rightarrow -\dfrac{K}{2}\) as \(n\rightarrow +\infty \).

Proof

It follows directly from Vieta’s formulas

$$\begin{aligned} x_n+y_n+z_n+\overline{z_n}=-a\mu _n^{\frac{1}{5}}. \end{aligned}$$
(2.1)

jointly with Lemma 2.1.

Remark 2.3

It will be useful later to know that for any \(\epsilon >0\) we have \(p(-a\mu _n^{\frac{1}{5}}-\epsilon )\) given by

$$\begin{aligned}{} & {} (a^2b-ac)\mu _n^{\frac{6}{5}}+(2ab-c)\epsilon \mu _n+(\epsilon ^2b)\mu _n^{\frac{4}{5}}+(a^3\epsilon )\mu _n^{\frac{3}{5}}\\{} & {} \quad +(3a^2\epsilon ^2+1)\mu _n^{\frac{2}{5}}+3a\epsilon ^3\mu _n^{\frac{1}{5}} +\epsilon ^4. \end{aligned}$$

2.1 Subcase \(ab=c\)

In this subcase the polynomial \(p_n\) becomes

$$\begin{aligned} p_n(\lambda )=\lambda ^4+a\mu _n^{\frac{1}{5}}\lambda ^3+b\mu _n^{\frac{4}{5}}\lambda ^2+ab\mu _n+\mu _n^{\frac{2}{5}}. \end{aligned}$$

We will prove that \(x_n+a\mu _n^{\frac{1}{5}}\rightarrow 0\) and consequently by Corollary 2.2 we have \(Re(z_n)\rightarrow 0\). From this, we will see that it is possible to prove a result of well-posedness for Eq. (1.1).

Proposition 1.4

If \(c>0\) and \(c=ab\), then \(Re(z_n)\rightarrow 0\). This means that, in the best case scenario, the problem (1.1) can only generate a strongly continuous semigroup on Y.

Proof

We will show that \(x_n\rightarrow -a\mu _n^{\frac{1}{5}}\) and the other claim follows from the Hille-Yosida Theorem, see e.g. [23, Chapter 1, Theorem 3.1]. Fixing \(\epsilon >0\) it is easy to see (Remark 2.3) that \(p(-a\mu _n^{\frac{1}{5}}+\epsilon )<0\) for large n. Since \(p_n(-a\mu _n^{\frac{1}{5}})>0\) we know that \(x_n\in (-a\mu _n^{\frac{1}{5}},-a\mu _n^{\frac{1}{5}}+\epsilon )\) for large n; that is, \(x_n\rightarrow -a\mu _n^{\frac{1}{5}}\) as \(n\rightarrow +\infty \).

Theorem 1.5

If \(c>0\) and \(c=ab\), then problem (1.1) is well-posedness on Y; that is, the linear operator associated with this problem is the infinitesimal generator of a strongly continuous semigroup on Y.

Proof

Define

$$\begin{aligned} z=u_t+aA^{\frac{1}{5}}u \end{aligned}$$
(2.2)

and

$$\begin{aligned} v=u_{ttt}+aA^{\frac{1}{5}}u_{tt}+bA^{\frac{4}{5}}u_t+abAu. \end{aligned}$$
(2.3)

Transferring our problem to the phase space

$$\begin{aligned} Z=X\times X\times X\times X \end{aligned}$$

endowed with the usual inner product, we define the vector \(W=\left[ {\begin{matrix} Au\\ Az\\ A^{\frac{3}{5}}z_t\\ A^{\frac{3}{5}}v \end{matrix}}\right] \). It is easy to see that (1.1) can be rewritten as a first-order abstract system of the form

$$\begin{aligned} W_t=\Lambda W,\quad t>0, \end{aligned}$$

where \(\Lambda \) denotes the unbounded linear operator given by

$$\begin{aligned} \Lambda =\begin{bmatrix} -aA^{\frac{1}{5}} &{}\quad 1 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad A^{\frac{2}{5}} &{}\quad 0 \\ 0&{}\quad -bA^{\frac{2}{5}} &{}\quad 0 &{}\quad 1\\ -I &{}\quad 0 &{}\quad 0 &{}\quad 0 \end{bmatrix}. \end{aligned}$$

Namely, this linear operator is closed and densely defined. Moreover, it is a maximally dissipative operator (see [23, Chapter 1, Definition 4.1]) on a space in which the norm is equivalent to the norm of space Z. This also implies that \(\left[ {\begin{matrix} u\\ u_t\\ u_{tt}\\ u_{ttt} \end{matrix}}\right] \in Y\).

2.2 Subcase \(c>ab\)

We will show that in this subcase the problem (1.1) is ill-posed, that is, the operator associated with the problem does not generate a strongly continuous semigroup of bounded linear operators.

Proposition 1.6

If \(c>0\) and \(c>ab\) then \(Re(z_n)\rightarrow +\infty \). Consequently, the unbounded linear operator associated with (1.1) does not generate a strongly continuous semigroup on Y.

Proof

It follows from Remark (2.3) that \(p_n(-a\mu _n^{\frac{1}{5}}-\mu _n^{\frac{1}{10}})<0\) for large n. Since \(p_n(\lambda )\rightarrow +\infty \) as \(\lambda \rightarrow -\infty \) we guarantee that \(x_n<-a\mu _n^{\frac{1}{5}}-\mu _n^{\frac{1}{10}}\). From (2.1) and Lemma 2.1 we obtain \(Re(z_n)\rightarrow +\infty \). This violates a necessary condition for the generation of a strongly continuous semigroup, see e.g. [23, Chapter 1, Corollary 3.8].

2.3 Subcase \(c<ab\)

Note that in this subcase we have to assume that \(a,b,c\ne 0\). We will see that problem (1.1) can be well-posed in the sense of the theory of semigroups of bounded linear operators. However, the semigroup associated with (1.1) cannot be analytic on Y.

Lemma 1.7

If \(a,b,c\ne 0\) and \(c<ab\) then \(x_n\rightarrow \dfrac{-c\mu _n^{\frac{1}{5}}}{b}\) as \(n\rightarrow +\infty \).

Proof

Firstly note that

$$\begin{aligned} p_n\left( \dfrac{-c\mu _n^{\frac{1}{5}}}{b}\right) =\dfrac{c^3(c-ab)}{b^4}\mu _n^{\frac{4}{5}}+\mu _n^{\frac{2}{5}} \end{aligned}$$

and therefore \(p_n(-cb^{-1}\mu _n^{\frac{1}{5}})<0\) for large n. Fixing \(\epsilon >0\) we can easily compute that

$$\begin{aligned} {\begin{matrix} p_n(-cb^{-1}\mu _n^{\frac{1}{5}}-\epsilon )&{}=\epsilon c\mu _n+(c^4b^{-4}-ac^3b^{-3}+\epsilon ^2b)\mu _n^{\frac{4}{5}}+(4c^3b^{-3}\epsilon -3ac^2b^{-2}\epsilon )\mu _n^{\frac{3}{5}}\\ &{}\quad +(1+6\epsilon ^2c^2b^{-2}-3\epsilon ^2acb^{-1})\mu _n^{\frac{2}{5}}+(4\epsilon ^3cb^{-1}-a\epsilon ^2)\mu _n^{\frac{1}{5}}+\epsilon ^4. \end{matrix}} \end{aligned}$$

Hence \(p_n(-cb^{-1}\mu _n^{\frac{1}{5}}-\epsilon )>0\) for large n which guarantees that

$$\begin{aligned} x_n\in (-cb^{-1}\mu _n^{\frac{1}{5}}-\epsilon ,-cb^{-1}\mu _n^{\frac{1}{5}}) \end{aligned}$$

for large n; that is, \(x_n\rightarrow -cb^{-1}\mu _n^{\frac{1}{5}}\) as \(n\rightarrow +\infty \).

Corollary 1.8

If \(a,b,c\ne 0\) and \(c<ab\) then \(Re(z_n)\rightarrow (c-ab)2^{-1}b^{-1}\mu _n^{\frac{1}{5}}\). Therefore the problem (1.1) can be well-posed in the sense of the theory of semigroups of bounded linear operators. Moreover, the solution of (1.1) cannot be analytic on Y in the sense of [11].

Proof

From Vieta’s formulas and Lemma 2.7 it follows immediately that \(Re(z_n)\rightarrow (c-ab)2^{-1}b^{-1}\mu _n^{\frac{1}{5}}\) and that

$$\begin{aligned} \left( \dfrac{Im(z_n)}{Re(z_n)}\right) ^2=\dfrac{4b^3\mu _n^{\frac{2}{5}}}{(c-ab)^2}\rightarrow +\infty , \end{aligned}$$

which concludes the proof.

3 Case \(c=0\)

3.1 Subcase \(a=b=c=0\)

In this case, the polynomial is given by

$$\begin{aligned} p_n(\lambda )=\lambda ^4+\mu _n^{\frac{2}{5}}=0. \end{aligned}$$

It is easy to see that the roots of \(p_n\) are \(\mu _n^{\frac{1}{10}}e^{\frac{\pi }{4}i}, \mu _n^{\frac{1}{10}}e^{\frac{3\pi }{4}i}, \mu _n^{\frac{1}{10}}e^{\frac{5\pi }{4}i}\) and \(\mu _n^{\frac{1}{10}}e^{\frac{7\pi }{4}i}\). To simplify our view, we can rewrite these roots as

$$\begin{aligned}{} & {} \mu _n^{\frac{1}{10}}\left( \dfrac{\sqrt{2}}{2}+i\dfrac{\sqrt{2}}{2}\right) , \mu _n^{\frac{1}{10}}\left( -\dfrac{\sqrt{2}}{2}+i\dfrac{\sqrt{2}}{2}\right) , \mu _n^{\frac{1}{10}}\left( -\dfrac{\sqrt{2}}{2}-i\dfrac{\sqrt{2}}{2}\right) \\{} & {} \quad \text {and}\ \mu _n^{\frac{1}{10}}\left( \dfrac{\sqrt{2}}{2}-i\dfrac{\sqrt{2}}{2}\right) . \end{aligned}$$

Theorem 1.9

If \(a=b=c=0\), then problem (1.1) is ill-posed in the sense of the theory of semigroups of bounded linear operators on Y.

Proof

Since \(\mu _n\rightarrow +\infty \) our next result follows directly from the Hille-Yosida Theorem, see e.g. [23, Chapter 1, Theorem 3.1].

3.2 Subcase \(a\ne 0\) and \(b=c=0\)

In this case we know that \(\Delta _n<0\) for large n and therefore our polynomial

$$\begin{aligned} \lambda ^4+a\mu _n^{\frac{1}{5}}\lambda ^3+\mu _n^{\frac{2}{5}} \end{aligned}$$

have two real roots \(x_n\), \(y_n\) and two complex roots \(z_n\) and \(\overline{z}_n\).

Lemma 1.10

If \(a\ne 0\) and \(b=c=0\) then \(x_n\rightarrow -a\mu _n^{\frac{1}{5}}\) as \(n\rightarrow +\infty \).

Proof

Given \(\epsilon >0\) we have

$$\begin{aligned} p_n(-a\mu _n^{\frac{1}{5}}+\epsilon )=-\epsilon a^3\mu _n^{\frac{3}{5}}+(3a^2\epsilon ^2+1)\mu _n^{\frac{2}{5}}-3a\epsilon ^3\mu _n^{\frac{1}{5}}+\epsilon ^4. \end{aligned}$$

Hence \(p_n(-a\mu _n^{\frac{1}{5}}+\epsilon )<0\) for large n. Since \(p_n(-a\mu _n^{\frac{1}{5}})>0\) for any n, we conclude that \(x_n\in (-a\mu _n^{\frac{1}{5}},-a\mu _n^{\frac{1}{5}}+\epsilon )\) for large n; that is, \(x_n\rightarrow -a\mu _n^{\frac{1}{5}}\) as \(n\rightarrow +\infty \).

Lemma 1.11

If \(a\ne 0\) and \(b=c=0\) then \(y_n\rightarrow - a^{-\frac{1}{3}}\mu _n^{\frac{1}{15}}\) as \(n\rightarrow +\infty \).

Proof

Fix \(\epsilon >0\) and note that

$$\begin{aligned} {\begin{matrix} p_n(-a^{-\frac{1}{3}}\mu _n^{\frac{1}{15}}-\epsilon )&{}= -3\epsilon a^{\frac{1}{3}}\mu _n^{\frac{5}{15}}+(a^{-\frac{4}{3}}-3\epsilon ^2a^{\frac{2}{3}})\mu _n^{\frac{4}{15}}+(4\epsilon a^{-1}-a\epsilon ^3)\mu _n^{\frac{3}{15}}\\ &{}\quad +6\epsilon ^2a^{-\frac{2}{3}}\mu _n^{\frac{2}{15}}+4\epsilon ^3a^{-\frac{1}{3}}\mu _n^{\frac{1}{15}}+\epsilon ^4. \end{matrix}} \end{aligned}$$

Thus \(p_n(-a^{-\frac{1}{3}}\mu _n^{\frac{1}{15}}-\epsilon )<0\) for large n. Since \(p_n(-a^{-\frac{1}{3}}\mu _n^{\frac{1}{15}})=a^{\frac{-4}{3}}\mu _n^{\frac{4}{15}}\), we guarantee that \(p_n(-a^{-\frac{1}{3}}\mu _n^{\frac{1}{15}})>0\) for large n and therefore

$$y_n\in (-a^{-\frac{1}{3}}\mu _n^{\frac{1}{15}}-\epsilon ,-a^{-\frac{1}{3}}\mu _n^{\frac{1}{15}})$$

when n is large. This means that \(|y_n+a^{-\frac{1}{3}}\mu _n^{\frac{1}{15}}|\rightarrow 0\) as \(n\rightarrow +\infty \).

Corollary 1.12

If \(a\ne 0\) and \(b=c=0\) then \(Re(z_n)\rightarrow +\infty \). Consequently, problem (1.1) is ill-posed in the sense of the theory of semigroups of bounded linear operators on Y.

3.3 Subscase \(b\ne 0\) and \(a=c=0\)

In this case we have the polynomial

$$\begin{aligned} \lambda ^4+b\mu _n^{\frac{4}{5}}\lambda ^2+\mu _n^{\frac{2}{5}} \end{aligned}$$
(3.1)

which has \(\Delta _n>0\) for large n and \(P_n=8b\mu _n^{\frac{4}{5}} >0\); that is, our polynomial has two pairs of non-real complex conjugate roots.

Lemma 1.13

If \(b\ne 0\) and \(a=c=0\) then all roots are imaginary for large n. Consequently, problem (1.1) can generate a strongly continuous semigroup of bounded linear operators on Y. Moreover, it cannot generate an analytic semigroup on Y in the sense of [11].

Proof

Denoting the four roots by \(z_n, \overline{z}_n, w_n, \overline{w}_n\) we know from Vieta’s formulas that

$$\begin{aligned} Re(w_n)=-Re(z_n), \ \ \ |z_n|^2|w_n|^2=\mu _n^{\frac{2}{5}} \end{aligned}$$
(3.2)

and

$$\begin{aligned} Re(w_n)(|z_n|^2-|w_n|^2)=0. \end{aligned}$$

Assume that \(Re(w_n)\ne 0\). Then

$$\begin{aligned} |z_n|^2=|w_n|^2=\mu _n^{\frac{1}{5}}. \end{aligned}$$
(3.3)

Writing \(w_n\) on its polar form \(w_n=r_ne^{i\theta _n}\) and using (3.1) it follows that

$$\begin{aligned} r_n^4e^{i4\theta _n}+b\mu _n^{\frac{4}{5}}r_n^2e^{i2\theta _n}+\mu _n^{\frac{2}{5}}=0 \end{aligned}$$

and therefore

$$\begin{aligned} r_n^2(r_n^2sin(4\theta _n)+b\mu _n^{\frac{4}{5}}sin(2\theta _n))=0. \end{aligned}$$
(3.4)

From (3.1) we have \(r_n=|w_n|\ne 0\) and this leads to

$$\begin{aligned} r_n^2sin(4\theta _n)+b\mu _n^{\frac{4}{5}}sin(2\theta _n)=0. \end{aligned}$$

Case \(sin(4\theta _n)=0\) we have \(sin(2\theta _n)=0\), and \(\theta _n=0\) or \(\theta _n=\frac{\pi }{2}\). From this, \(sin(4\theta _n)\ne 0\); that is,

$$\begin{aligned} r_n^2=\dfrac{- sin(2\theta _n)b\mu _n^{\frac{4}{5}}}{sin(4\theta _n)}. \end{aligned}$$
(3.5)

Putting (3.3) and (3.5) together we obtain

$$\begin{aligned} -sin(2\theta _n)b\mu _n^{\frac{3}{5}}=sin(4\theta _n). \end{aligned}$$

Thus

$$\begin{aligned} sin(2\theta _n)(b\mu _n^{\frac{3}{5}}+2cos(2\theta _n))=0. \end{aligned}$$

Since \(b\mu _n^{\frac{3}{5}}+2cos(2\theta _n)\ne 0\) for large n we conclude that \(sin(2\theta _n)=0\) and

$$\begin{aligned} \theta _n=0\ \ \text {or}\ \ \theta _n=\dfrac{\pi }{2} \end{aligned}$$
(3.6)

which contradicts our assumption that \(w_n\notin \mathbb {R}\) and \(Re(w_n)\ne 0\). Therefore

$$Re(w_n)=Re(z_n)=0$$

for large n we conclude the proof.

Theorem 1.14

If \(b\ne 0\) and \(a=c=0\) then the problem (1.1) generates a strongly continuous semigroup on Y.

Proof

It follows similarly to Theorem 2.5.

3.4 Subcase \(a,b\ne 0\) and \(c=0\)

In this case we have the polynomial

$$\begin{aligned} \lambda ^4+a\mu _n^{\frac{1}{5}}\lambda ^3+b\mu _n^{\frac{4}{5}}\lambda ^2+\mu _n^{\frac{2}{5}} \end{aligned}$$
(3.7)

which has \(\Delta _n>0\) for large n. Moreover, \(P_n=8b\mu _n^{\frac{4}{5}}-3a^2\mu _n^{\frac{2}{5}}>0\) for large n, which implies that our polynomial has four complex roots \(z_n, \overline{z_n}, w_n\) and \(\overline{w_n}\).

Lemma 1.15

Two of the four roots of polynomial (3.7); namely \(w_n\) and \(\overline{w_n}\) satisfy

$$|w_n|=| \overline{w_n}|\rightarrow b^{-\frac{1}{2}}\mu _n^{-\frac{1}{5}}\ \ \text {as} \ \ n\rightarrow +\infty .$$

Proof

From Vieta’s formulas, we can easily obtain that

$$\begin{aligned}{} & {} |w_n|^2|z_n|^2=\mu _n^{\frac{2}{5}}, \end{aligned}$$
(3.8)
$$\begin{aligned}{} & {} Re(z_n)|w_n|^2=-Re(w_n)| z_n|^2 \end{aligned}$$
(3.9)
$$\begin{aligned}{} & {} |z_n|^2+|w_n|^2+4Re(z_n)Re(w_n)=b\mu _n^{\frac{4}{5}} \end{aligned}$$
(3.10)

and

$$\begin{aligned} Re(w_n)+Re(z_n)=\dfrac{-a\mu _n^{\frac{1}{5}}}{2}. \end{aligned}$$
(3.11)

Since \(w_n,z_n\ne 0\) we can put (3.8) and (3.9) into Eq. (3.10) to obtain that

$$\begin{aligned} \dfrac{\mu _n^{\frac{2}{5}}}{|w_n|^2}+|w_n|^2-\dfrac{4Re(w_n)^2\mu _n^{\frac{2}{5}}}{|w_n|^4}=b\mu _n^{\frac{4}{5}}. \end{aligned}$$

which leads us to

$$\begin{aligned} \mu _n^{\frac{2}{5}}|w_n|^2+|w_n|^6-4Re(w_n)^2\mu _n^{\frac{2}{5}}-b\mu _n^{\frac{4}{5}}|w_n|^4=0. \end{aligned}$$
(3.12)

Now, putting (3.9) and (3.11) together we have

$$\begin{aligned} \dfrac{-a\mu _n^{\frac{1}{5}}|w_n|^2 }{2}=Re(w_n)(|w_n|^2-|z_n|^2). \end{aligned}$$
(3.13)

Since \(w_n\ne 0\) we know from (3.13) that \(|w_n|^2-|z_n|^2\ne 0\) and consequently

$$\begin{aligned} Re(w_n)=\frac{-a\mu _n^{\frac{1}{5}}|w_n|^2}{2(|w_n|^2-|z_n|^2)}. \end{aligned}$$

Putting this equation together with (3.8) we have (the well-defined) equation

$$\begin{aligned} Re(w_n)=\dfrac{-a\mu _n^{\frac{1}{5}}|w_n|^4}{2(|w_n|^4-\mu _n^{\frac{2}{5}})}. \end{aligned}$$
(3.14)

Finally, putting (3.12) and (3.14) together we obtain

$$\begin{aligned} \mu _n^{\frac{2}{5}}|w_n|^2+|w_n|^6-\dfrac{a^2\mu _n^{\frac{4}{5}}|w_n| ^8}{|w_n| ^8-2\mu _n^{\frac{2}{5}}|w_n| ^4+\mu _n^{\frac{4}{5}}}-b\mu _n^{\frac{4}{5}}|w_n|^4=0. \end{aligned}$$

From this we have

$$\begin{aligned} (\mu _n^{\frac{2}{5}}|w_n|^2+|w_n|^6-b\mu _n^{\frac{4}{5}}|w_n|^4)(|w_n| ^8-2\mu _n^{\frac{2}{5}}|w_n| ^4+\mu _n^{\frac{4}{5}})-a^2\mu _n^{\frac{4}{5}}|w_n| ^8=0 \end{aligned}$$

and consequently

$$\begin{aligned} |w_n|^{14}-b\mu _n^{\frac{4}{5}}|w_n|^{12}-\mu _n^{\frac{2}{5}}\parallel&w_n\parallel ^{10}+(2b\mu _n^{\frac{6}{5}}-a^2\mu _n^{\frac{4}{5}})|w_n|^8\\&-\mu _n^{\frac{4}{5}}|w_n|^6-b\mu _n^{\frac{8}{5}}|w_n|^4+\mu _n^{\frac{6}{5}}|w_n|^2=0. \end{aligned}$$

Dividing both sides by \(|w_n|^2\) we finally have

$$\begin{aligned}&|w_n|^{12}-b\mu _n^{\frac{4}{5}}|w_n|^{10}-\mu _n^{\frac{2}{5}}|w_n|^{8}\\&\quad +(2b\mu _n^{\frac{6}{5}}-a^2\mu _n^{\frac{4}{5}})|w_n|^6-\mu _n^{\frac{4}{5}}|w_n|^4-b\mu _n^{\frac{8}{5}}|w_n|^2+\mu _n^{\frac{6}{5}}=0. \end{aligned}$$

Define the polynomial \(q_n:\mathbb {R}\rightarrow \mathbb {R}\) as

$$\begin{aligned} q_n(x)=x^{12}-b\mu _n^{\frac{4}{5}}x^{10}-\mu _n^{\frac{2}{5}}x^{8}+(2b\mu _n^{\frac{6}{5}}-a^2\mu _n^{\frac{4}{5}})x^6-\mu _n^{\frac{4}{5}}x^4-b\mu _n^{\frac{8}{5}}x^2+\mu _n^{\frac{6}{5}}. \end{aligned}$$

From

$$\begin{aligned} q_n(b^{-\frac{1}{2}}\mu _n^{-\frac{1}{5}})=b^{-6}\mu _n^{-\frac{12}{5}}-2b^{-4}\mu _n^{-\frac{6}{5}}-a^2b^{-3}\mu _n^{-\frac{2}{5}}-2b^{-2} \end{aligned}$$

we obtain that \(q_n(b^{-\frac{1}{2}}\mu _n^{-\frac{1}{5}})<0\) for large n. It is easy to see that for any \(\epsilon >0\) we have \(q_n(b^{-\frac{1}{2}}\mu _n^{-\frac{1}{5}}-\epsilon )>0\) for large n. Thus, given \(\epsilon >0\) we have

$$|w_n|=| \overline{w_n}|\in (b^{-\frac{1}{2}}\mu _n^{-\frac{1}{5}}-\epsilon ,b^{-\frac{1}{2}}\mu _n^{-\frac{1}{5}})$$

for large n, which completes the proof.

Lemma 1.16

Assume that \(w_n\) is a root of polynomial (3.7) such that

$$| w_n| \rightarrow b^{-\frac{1}{2}}\mu _n^{-\frac{1}{5}}$$

as \(n\rightarrow +\infty \). Then \(\dfrac{Re(w_n)}{Im(w_n)}\rightarrow 0\) as \(n\rightarrow +\infty \).

Proof

We will denote the polar form of \(w_n\) as \(r_ne^{i\theta _n}\). Remember that

$$r_n\rightarrow b^{-\frac{1}{2}}\mu _n^{-\frac{1}{5}}\ \ \text {as}\ \ n\rightarrow +\infty .$$

We just need to prove that \(\theta _n\rightarrow \dfrac{\pi }{2}\) or \(\theta _n\rightarrow \dfrac{3\pi }{2}.\)

Since \(w_n\) is a root of (3.7), we have

$$\begin{aligned} r_n^2(r_n^2cos(4\theta _n)+a\mu _n^{\frac{1}{5}}r_ncos(3\theta _n)+b\mu _n^{\frac{4}{5}}cos(2\theta _n))=-\mu _n^{\frac{2}{5}}. \end{aligned}$$

It is easy to see that a root from (3.7) must be different from zero. Thus

$$\begin{aligned} r_n^2cos(4\theta _n)+a\mu _n^{\frac{1}{5}}r_ncos(3\theta _n)+b\mu _n^{\frac{4}{5}}cos(2\theta _n)=-\mu _n^{\frac{2}{5}}r_n^{-2}. \end{aligned}$$

Since \(-\mu _n^{\frac{2}{5}}r_n^{-2}\rightarrow -b\mu _n^{\frac{4}{5}}\) we conclude that

$$\begin{aligned} r_n^2cos(4\theta _n)+a\mu _n^{\frac{1}{5}}r_ncos(3\theta _n)+b\mu _n^{\frac{4}{5}}cos(2\theta _n)\rightarrow -b\mu _n^{\frac{4}{5}}. \end{aligned}$$

From the fact that \(r_n\rightarrow b^{-\frac{1}{2}}\mu _n^{-\frac{1}{5}}\) it is easy to see that \(\theta _n\rightarrow \dfrac{\pi }{2}\) or \(\theta _n\rightarrow \dfrac{3\pi }{2}\) as \(n\rightarrow +\infty \).

Corollary 1.17

There exist two roots of polynomial (3.7); namely \(w_n\) and \(\overline{w_n}\) such that \(w_n\rightarrow b^{-\frac{1}{2}}\mu _n^{-\frac{1}{5}}i\) and \(\overline{w_n}\rightarrow -b^{-\frac{1}{2}}\mu _n^{-\frac{1}{5}}i\) as \(n\rightarrow +\infty \).

Proof

It follows immediately from Lemmas 3.7 and 3.8.

Corollary 1.18

If \(a,b\ne 0\) and \(c=0\) then the problem (1.1) can generate a strongly continuous semigroup. Furthermore, the semigroup cannot be of an analytic type on Y in the sense of [11].

Proof

From Corollary 3.9 and Vieta’s formulas, we have that

$$\begin{aligned} Re(z_n)\rightarrow -\dfrac{a}{2}\mu _n^{\frac{1}{5}}\ \text {as}\ n\rightarrow +\infty . \end{aligned}$$

This means that problem (1.1) can generate a strongly continuous semigroup. Once again from Vieta’s formulas we can obtain that

$$\begin{aligned} (Im(z_n))^2\rightarrow b\mu _n^{\frac{4}{5}}-\dfrac{a^2}{4}\mu _n^{\frac{2}{5}}\ \text {as}\ n\rightarrow +\infty , \end{aligned}$$

implying that

$$\begin{aligned} \dfrac{(Im(z_n))^2}{(Re(z_n))^2}\rightarrow +\infty \ \text {as}\ n\rightarrow +\infty . \end{aligned}$$

Therefore, problem (1.1) cannot have an analytic semigroup on Y as a solution operator.