Skip to content
BY 4.0 license Open Access Published by De Gruyter February 9, 2022

Optimal decay rate for higher–order derivatives of solution to the 3D compressible quantum magnetohydrodynamic model

  • Juan Wang and Yinghui Zhang EMAIL logo

Abstract

We investigate optimal decay rates for higher–order spatial derivatives of strong solutions to the 3D Cauchy problem of the compressible viscous quantum magnetohydrodynamic model in the H5 × H4 × H4 framework, and the main novelty of this work is three–fold: First, we show that fourth order spatial derivative of the solution converges to zero at the L2-rate(1+t)-114 , which is same as one of the heat equation, and particularly faster than the L2-rate(1+t)-54 in Pu–Xu [Z. Angew. Math. Phys., 68:1, 2017] and the L2-rate(1+t)-94 , in Xi–Pu–Guo [Z. Angew. Math. Phys., 70:1, 2019]. Second, we prove that fifth–order spatial derivative of density ρ converges to zero at the L2-rate(1+t)-134 , which is same as that of the heat equation, and particularly faster than ones of Pu–Xu [Z. Angew. Math. Phys., 68:1, 2017] and Xi–Pu–Guo [Z. Angew. Math. Phys., 70:1, 2019]. Third, we show that the high-frequency part of the fourth order spatial derivatives of the velocity u and magnetic B converge to zero at the L2-rate(1+t)-134 , which are faster than ones of themselves, and totally new as compared to Pu–Xu [Z. Angew. Math. Phys., 68:1, 2017] and Xi–Pu–Guo [Z. Angew. Math. Phys., 70:1, 2019].

MSC 2010: 35B40; 35Q35; 76W05

1 Introduction

In this paper, we consider optimal decay rates for higher–order spatial derivatives of strong solutions to the following 3D compressible viscous quantum magnetohydrodynamic (vQMHD) model:

(1.1) {ρt+div(ρu)=0,(ρu)t+div(ρuu)-μΔu-(μ+λ)divu+P(ρ)-ϑ22ρ(Δρρ)=(×B)×B,Bt-×(u×B)=-×(ν×B),divB=0,

where t ≥ 0 is time and x ∈ ℝ3 is the spatial coordinate, and the symbol ⊗ is the Kronecker tensor product. The unknown functions ρ = ρ(x, t) is the density, u = (u1, u2, u3)(x, t) denotes the velocity, and B = (B1, B2, B3)(x, t) represents the magnetic field. P = P(ρ) = γ (a > 0, γ ≥ 1) stands for the pressure. The constant viscosity coefficients μ and λ satisfy the physical conditions: μ > 0, 23μ+λ0 , and ν > 0 denotes the magnetic diffusion coefficient. The constant ϑ > 0 represents the Planck constant. The expression Δρρ is the so-called Bohm quantum potential satisfying:

2ρ(Δρρ)=div(ρ2ρ)=Δρ+|ρ|2ρρ2-ρΔρρ-ρ2ρρ=Δρ-4div(ρρ).

The system (1.1) is supplemented with the following initial data

(1.2) (ρ,u,B)|t=0=(ρ0(x),u0(x),B0(x)).

Moreover, we assume that when the space variable goes to infinity, the initial perturbation satisfies

(1.3) lim|x|(ρ0-1,u0,B0)(x)=0.

1.1 History of the problem

The quantum fluid model can provide many pieces of information for the particles in the semiconductor simulation, and it could be used to describe quantum semiconductors [6], weakly interacting Bose gases [8], and quantum trajectories of Bohmian mechanics [34]. The quantum magnetohydrodynamic (QMHD) model plays an important role in modeling and simulating electron transport, which was extended by Hass [10] later from a Wigner–Maxwell system. Moreover, this model could be used to describe global properties of quantum plasmas. It is worth mentioning that system (1.1) will reduce to the compressible MHD equations without quantum effects. There is a vast literature addressing the decay and other asymptotic behaviors of compressible fluid equations, we refer to previous studies [1, 2, 3, 4, 5, 11, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 24, 25, 29, 30, 31, 32] and references therein.

In what follows, we only review some results closely related for simplicity. The study for decay rates of solutions to the QMHD model has attracted much attention of mathematicians. Pu and Guo [25] established the optimal decay rates of classical solutions near constant states via the spectral method in ℝ3. In [27], Pu and Xu showed that the classical solution of the vQMHD model has the following decay rate:

k(ρ-1)(t)H5-k+ku(t)H4-k+kB(t)H4-kC(1+t)-3+2k4,k=0,1.

Pu–Xu [28] employed the pure energy method developed by Guo–Wang [9] to obtain the optimal decay rates of higher-order spatial derivatives of solutions to the full hydrodynamic equations with quantum effects under the condition that the initial perturbation belongs to (HN+2H˙-s)×(HN+1H˙-s)×(HNH˙-s) for N ≥ 3 and s[0,32) . Recently, by making full use of Fourier splitting method, Xi–Pu–Guo [35] improved the work of Pu–Xu [27], and particularly they got

(1.4) k(ρ-1)(t)H5-k+ku(t)H4-k+kB(t)H4-kC(1+t)-3+2k4,k=0,1,2,3.

For more results about the large time behavior of solutions to the quantum fluid model, readers can refer to [26, 36] and references therein.

It is clear that decay rate of k(k = 0, 1, 2, 3) order spatial derivative of solution in (1.4) is optimal in the sense that it coincides with decay rate of solution to the heat equation. However, decay rate in (1.4) implies that fourth order spatial derivative of solution converges to zero at L2-rate(1+t)-94 , which is slower than the L2-rate(1+t)-114 of solution to the heat equation. Moreover, fifth order spatial derivative of density ρ converges to zero at L2-rate(1+t)-94 , which is slower than the L2-rate(1+t)-134 of solution to the heat equation. Therefore, decay rates of fourth order spatial derivative of solution and fifth order spatial derivative of density ρ are not optimal in this sense. The main purpose of this paper is to give a clear answer to this issue. More precisely, our main results can be outlined as follows. Firstly, we show that fourth order spatial derivative of the solution converges to zero at the L2-rate(1+t)-114 . Secondly, we prove that fifth–order spatial derivative of density ρ converges to zero at the L2-rate(1+t)-134 . Thirdly, we show that the high-frequency part of the fourth order spatial derivatives of velocity u and magnetic B converge to zero at the L2-rate(1+t)-134 , which are faster than ones of themselves, and totally new as compared to Pu–Xu [27] and Xi–Pu–Guo [35].

1.2 Notation

In this paper, we use Hk(ℝ3) to denote the usual Sobolev spaces with norm ‖ · ‖Hk. Generally, we use Lp, 1 ≤ p ≤ ∞ to denote the usual Lp (ℝ3) spaces with norm ‖ · ‖Lp. The notation ab means that aCb, where the universal constant C > 0 may be from line to line but independent of time t. Similarly, the notation ab means that aCb for a universal positive constant which is independent of time t. We define

ku=xαuiα=k,i=1,2,3,u=(u1,u2,u3).

For a radial function ϕC0(ξ3) such that ϕ(ξ) = 1 when |ξ | ≤ 1 and ϕ(ξ ) = 0 when |ξ | ≥ 2, we define the low–frequency part and the high–frequency part of f as follows

fl=𝔉-1[ϕ(ξ)f^], and fh=𝔉-1[(1-ϕ(ξ))f^].

1.3 Main results

Before stating our main results, let us recall the following results obtained in previous studies [28, 35], which will be used in this paper frequently.

Theorem 1.1

(see [28, 35]) Suppose that the initial data (ρ0 − 1, u0, B0) ∈ H5 × H4 × H4. There exists a small constant δ > 0 such that if

(1.5) ρ0-1H5+u0H4+B0H4δ,

then the solution (ρ, u, B) of (1.1)(1.3) satisfy for all T ≥ 0

(1.6) (ρ-1,u,B)(t)H42+ϑρ(t)H42+0t(u,B,ϑρ)(s)H42dsC(ρ0-1H52+u0H42+B0H42).

Moreover, provided that ‖;(ρ0 − 1, u0, B0)‖L1 is finite additionally, then the solution (ρ, u, B) satisfy

(1.7) k(ρ-1)(t)H5-k+ku(t)H4-k+kB(t)H4-kC(1+t)-3+2k4,

with k = 0, 1, 2, 3. Here, the positive constant C is independent of time.

Our main purpose in this paper is to establish the upper optimal decay rates for fourth order spatial derivative of the solution and fifth order spatial derivatives of the density ρ which are the same as those of the heat equation. Our main results are stated in the following theorems:

Theorem 1.2

Under all the assumptions in Theorem 1.1, the global solution (ρ, u, B) has the following time decay rate for all tT

(1.8) 4(ρ-1)(t)H1+4u(t)L2+4B(t)L2(1+t)-114.

Here T is a positive large time.

Theorem 1.3

Under all the assumptions in Theorem 1.1, the global solution (ρ, u, B) has the following time decay rate for all tT

(1.9) 5(ρ-1)(t)L2+4uh(t)L2+4Bh(t)L2(1+t)-134.

Here T is a positive large time.

Remark 1.4

It is interesting to make a comparison between Theorem 1.2 and Theorem 1.1, where the authors derived decay rate of fourth order spatial derivative of solution is the same as one of third order spatial derivative. We show that fourth order spatial derivative of the solution converges to zero at the L2-rate(1+t)-114 , which is same as that of the heat equation, and particularly faster than the L2-rate(1+t)-54 in Pu–Xu [27] and the L2-rate(1+t)-94 in Xi–Pu–Guo [35]. Therefore, our decay rate is optimal, and improve the results of [28, 35].

Remark 1.5

Compared to the main results (1.7) in [28, 35], decay rate in (1.9) is totally new and gives optimal decay rate of fifth order spatial derivative of density ϱ, which is same as that of the heat equation. In addition, it also gives optimal decay rates of the high-frequency part of fourth order spatial derivatives of the velocity u and magnetic B, which are faster than ones of themselves.

Now, let’s sketch the strategy of proving Theorem 1.2Theorem 1.3 and explain some main difficulties and techniques involved in the process. Roughly speaking, our methods mainly involve low–frequency and high–frequency decomposition technique and delicate energy estimates. Our strategy can be outlined as follows.

For the proof of Theorem 1.2, we hope to establish the optimal decay rate for fourth order spatial derivatives of solution for the compressible viscous quantum magnetohydrodynamic equations (1.1). For the convenience of calculation, we linearize the system (1.1) around the equilibrium state, and then rewrite the linearized system in terms of the variables ϱ, u, B, which is stated in (2.1). We only need to make full use of the benefit of low-frequency and high-frequency decomposition to deal with high-frequency part of ∇4(ϱ, u, B). The proof mainly involves the following three steps. First, we derive high–frequency L2 energy estimate of fourth order spatial derivative of the solution to the compressible vQMHD model:

12ddt3|4ϱh|2+|4uh|2+|4Bh|2+ϑ24|5ϱh|2dx+(2μ+λ)3|5uh|2dx+ν3|5Bh|2dx(1+t)-112+(1+t)-2(4ϱhH12+5BhL22)+(1+t)-32(5uhL22+6ϱhL22).

Second, by noticing that the above energy inequality only involves dissipative estimates of uh and Bh. In order to explore the dissipative estimate of ϱh, the main idea here is to employ the new interactive energy functional to get

ddt34uh5ϱhdx+125ϱhL22+ϑ246ϱhL22(1+t)-112+5uhL22+(1+t)-32(6ϱhL22+5uhL22).

Last, we choose two sufficiently large positive constants D0 and T1, and define the temporal energy functional

(t)=D0(4(ϱ,u,B)hL22+ϑ225ϱhL22)+34uh5ϱhdx,

for t ≥ 0, where it is mentioned that ℰ(t) is equivalent to 4ϱhH12+4(u,B)hL22 . Combining the above estimates, we can obtain

ddt(t)+4ϱh(t)H12+4uh(t)L22+4Bh(t)L22(1+t)-112,

which together with Gronwall’s argument and low-frequency decay rate gives rise to (1.8), and thus completes the proof of Theorem 1.2.

For the proof of Theorem 1.3, we hope to establish the fifth order spatial derivatives of the density ρ. Our main idea is to find the optimal decay rate at low–frequency and high–frequency of the fifth order spatial derivative of density, and then use the properties of high-frequency and low-frequency decomposition to obtain the optimal decay rates of the fifth order spatial derivative of density. For the low-frequency part ‖∇5 ϱlL2, by using the equation (2.1)1, Theorem 1.2, Duhamel’s principle, Plancherel theorem, Hölder’s inequality and Hausdorff–Young’s inequality, we have

5ϱl(t)L2 (1+t)-134(ϱ,u,B)(0)L1+0t2(1+t-τ)-134G(τ)L1dτ+t2t(1+t-τ)-54|ξ|4G^l(τ)Ldτ.

After some detailed calculations, we can get

(1.10) 5ϱlL2(1+t)-134.

For the high-frequency part ‖∇5 ϱlL2, it is easy to find that the calculation in this part is similar to that of section 2, we only need to replace the decay rate of fourth order space derivative from (1+t)-94 to (1+t)-114 . We finally arrive at

(1.11) 4ϱh(t)H1+4uh(t)L2+4Bh(t)L2(1+t)-134.

Combining the estimates (1.10) and (1.11), we conclude the proof of Theorem 1.3.

2 Proof of Theorem 1.2

In this section, we will devote ourselves to proving Theorem 1.2. We suppose that all the conditions of Theorem 1.1 are in force in this and next section. Without loss of generality, we assume that P′(1) = 1. Let us setting ϱ = ρ − 1, we can rewrite the system (1.1) in the perturbation form as

(2.1) {ϱt+divu=G1,ut-μΔu-(μ+λ)divu+ϱ-ϑ24Δϱ=G2,Bt-νΔB=G3.

Here the nonlinear source terms G1, G2 and G3 are given by

G1=-ϱdivu-uϱ,

G2=-uu+ϱ(ϱ+1)(μΔu+(μ+λ)divu)-(P(ϱ+1)ϱ+1-1)ϱ-ϑ24ϱϱ+1Δϱ+ϑ24(|ϱ|2ϱ(1+ϱ)3-ϱΔϱ(1+ϱ)2-ϱ2ϱ(1+ϱ)2)+11+ϱ((×B)×B),G3=×(u×B).

The initial data are given as

(2.2) (ϱ,u,B)(x,t)|t=0=(ϱ0,u0,B0)(x)(0,0,0) as |x|.

First, we recall the L2 time decay rates on the linearized system for the first two equations (2.1).

Lemma 2.1

(see [27]) Let s ≥ 0 be an integer. Assume that (ϱ, u) is the solution of the linearized system for the first two equations in (2.1) with initial data ϱHs+1L1, u0HsL1. Then, for any t ≥ 0, it holds that

ϱ(t)L2C(1+t)-34((ϱ,u0)L1+(ϱ0,u0)L2),

(2.3) k+1ϱ(t)L2C(1+t)-34-k+12((ϱ0,u0)L1+(k+1ϱ,ku0)L2),

(2.4) ku(t)L2C(1+t)-34-k2((ϱ0,u0)L1+(k+1ϱ,ku0)L2),

for 0 ≤ ks.

Next, we recall the L2 time decay rate on the linearized system for the third equation in (2.1).

Lemma 2.2

(see [33]) Let s ≥ 0 be an integer. Assume that B is the solution of the linearized system for the third equation in (2.1) with initial data B0HsL1. Then, for any t ≥ 0, it holds that

(2.5) kBL2C(1+t)-34-k2B0L1,

for 0 ≤ ks.

Finally, we deduce the L2 time decay rate on the nonlinear system (2.1).

Lemma 2.3

Assume that the assumptions of Theorem 1.1 are in force, the solution (ϱ, u, B) of the nonlinear system (2.1) satisfies the following decay estimate:

(2.6) 4(ϱl,ul,Bl)(t)L2(1+t)-114.

Proof

By virtue of the equation (2.1), Lemma 2.1, Lemma 2.2, Duhamel’s principle, Plancherel theorem, Hölder’s inequality and Hausdorff–Young’s inequality, we have

(2.7) 4(ϱl,ul,Bl)(t)L2 (1+t)-114(ϱ,u,B)(0)L1+0t2(1+t-τ)-114G(τ)L1dτ+t2t(1+t-τ)-54|ξ|3G^l(τ)Ldτ.

Next, we shall estimate the second term and the third term on the right-hand side of (2.7). To begin with, by virtue of Lemma 4.2, we can bound the term by

(2.8) G(τ)L1div(ϱu)(τ)L1+uu(τ)L1+ϱϱ+1[μΔu+(μ+λ)divu](τ)L1+[P(ϱ+1)ϱ+1-P(1)]ϱ(τ)L1+ϑ24ϱϱ+1Δϱ(τ)L1+ϑ24(|ϱ|2ϱ(1+ϱ)3-ϱΔϱ(1+ϱ)2-ϱ2ϱ(1+ϱ)2)(τ)L1+11+ϱ((×B)×B)(τ)L1+×(u×B)(τ)L1(ϱ,u)(τ)L2(ϱ,u)(τ)L2+ϱ(τ)L22u(τ)L2+ϱ(τ)L23ϱ(τ)L2+ϱ(τ)L22ϱ(τ)L2+B(τ)L22B(τ)L2+(u,B)(τ)L2(u,B)(τ)L2(1+t)-32,

and

(2.9) |ξ|3G^l(τ)L3Gl(τ)L13(div(ϱu))(τ)L1+3(uu)(τ)L1+2(ϱϱ+1[μΔu+(μ+λ)divu])(τ)L1+3([P(ϱ+1)ϱ+1-P(1)]ϱ)(τ)L1+2(ϑ24ϱϱ+1Δϱ)(τ)L1+3(ϑ24(|ϱ|2ϱ(1+ϱ)3-ϱΔϱ(1+ϱ)2-ϱ2ϱ(1+ϱ)2))(τ)L1+3(11+ϱ((×B)×B))(τ)L1+3(×(u×B)(τ))L1(ϱ,u)(τ)L24(ϱ,u)(τ)L2+2u(τ)L22+ϱ(τ)L25ϱ(τ)L2+ϱ(τ)L25ϱ(τ)L2+ϱ(τ)L24ϱ(τ)L2+2ϱ(τ)L23ϱ(τ)L2+B(τ)L23B(τ)L2+(u,B)(τ)L24(u,B)(τ)L2(1+t)-3.

Substituting (2.8)(2.9) into (2.7) gives (2.6).

Proof of Theorem 1.2

The first step is to establish the energy estimate for the fourth order spatial derivative of solution. Taking

𝔉-1(1-ϕ(ξ))4(2.1)1,4ρh+𝔉-1(1-ϕ(ξ))4(2.1)2,4uh+𝔉-1(1-ϕ(ξ))4(2.1)3,4Bh,

and using integration by parts, we can obtain

(2.10) 12ddt3|4ϱh|2+|4uh|2+|4Bh|2+ϑ24|5ϱh|2dx+(2μ+λ)3|5uh|2dx+ν3|5Bh|2dx=4G1h,4ϱh+4G2h,4uh+4G3h,4Bh+h245G1h,5ϱh.

The right-hand side of the above equation can be estimated as follows. Firstly, we have

(2.11) 4G1h,4ϱh=-4(ϱdivu+uϱ)h,4ϱh=-4(ϱdivu)h,4ϱh-4(uϱ)h,4ϱh:=I1+I2.

For term I1, by using Hölder’s inequality, Lemma 4.2, Lemma 4.3, Lemma 4.1, Young’s inequality and Sobolev interpolation theorem, we arrive at

(2.12) |I1|3(ϱdivu)hL25ϱhL23(ϱdivu)L25ϱhL2(ϱL4uL2+uL33ϱL6)5ϱhL2(ϱL2122ϱL2124uL2+uL2122ϱL2124ϱL2)5ϱhL2(1+t)-1545ϱhL2(1+t)-112+(1+t)-25ϱhL22.

For term I2, we can rewrite it as follows

(2.13) I2=-4(uϱ)h,4ϱh=-4(uϱ)+4(uϱ)l,4ϱh=-4(uϱh)-4(uϱl)+4(uϱ)l,4ϱh=-4(uϱh),4ϱh+-4(uϱl),4ϱh+4(uϱ)l,4ϱh:=I2,1+I2,2+I2,3.

By routine checking, one may show that

4(uϱh)=u5ϱh+4u4ϱh+62u3ϱh+43u2ϱh+44uϱh.

The integration by part yields directly

3u(4ϱh)4ϱhdx=-123divu|4ϱh|2dx,

and hence, we show that

(2.14) |divu4ϱh,4ϱh|divuL4ϱhL22divuL4ϱhL222uL2123uL2124ϱhL22(1+t)-1744ϱhL2(1+t)-112+(1+t)-34ϱhL22.

Similar to the estimate (2.14), we have

(2.15) |u4ϱh,4ϱh|(1+t)-112+(1+t)-34ϱhL22,

and

(2.16) |-62u3ϱh,4ϱh|2uL33ϱhL64ϱhL22uL33ϱL64ϱhL22uL2123uL2124ϱL24ϱhL2(1+t)-1744ϱhL2(1+t)-112+(1+t)-34ϱhL22,

(2.17) |-43u2ϱh,4ϱh|2ϱhL33uL64ϱhL22ϱL33uL64ϱhL22ϱL2123ϱL2124uL24ϱhL2(1+t)-1744ϱhL2(1+t)-112+(1+t)-34ϱhL22,

(2.18) |-4uϱh,4ϱh|ϱhL4uL24ϱhL22ϱL2123ϱL2124uL24ϱhL2(1+t)-1744ϱhL2(1+t)-112+(1+t)-34ϱhL22,

and hence, it follows that

(2.19) |I2,1|(1+t)-112+(1+t)-34ϱhL22.

For the term I2,2, we have

(2.20) |I2,2|4(uϱl)L24ϱhL2(uL5ϱlL2+ϱlL4uL2)4ϱhL2(uL2122uL2124ϱL2+2ϱlL2123ϱlL2124uL2)4ϱhL2(uL2122uL2124ϱL2+ϱL2122ϱL2124uL2)4ϱhL2(1+t)-1544ϱhL2(1+t)-112+(1+t)-24ϱhL22.

For the term I2,3, we get

(2.21) |I2,3|4(uϱ)lL24ϱhL23(uϱ)L24ϱhL2(uL4ϱL2+ϱL33uL6)4ϱhL2(uL2122uL2124ϱL2+ϱL2122ϱL2124uL2)4ϱhL2(1+t)-1544ϱhL2(1+t)-112+(1+t)-24ϱhL22.

Summing up estimates (2.19)(2.21), we arrive at

(2.22) |I2|(1+t)-112+(1+t)-24ϱhH12.

Applying equation (2.1), it holds that

(2.23) 4G2h,4uh=-4(uu)h,4uh+4{ϱ(ϱ+1)[μΔv+(μ+λ)divu]}h,4uh+-4{[P(ϱ+1)ϱ+1-P(1)]ϱ}h,4uh+-4(ϑ24ϱϱ+1Δϱ)h,4uh+4[ϑ24(|ϱ|2ϱ(1+ϱ)3-ϱΔϱ(1+ϱ)2-ϱ2ϱ(1+ϱ)2)]h,4uh+4[11+ϱ((×B)×B)]h,4uh:=j=38Ij.

The first term on the right-hand side of the above equation can be estimated as follows

(2.24) |I3|3(uu)hL25uhL23(uu)L25uhL2(uL4uL2+3uL6uL3)5uhL2uL2122uL2124uL25uhL2(1+t)-1545uhL2(1+t)-112+(1+t)-25uhL22.

Similarly, using the integration by parts, the terms I4I8 can be estimated as follows

(2.25) |I4|3{ϱ(ϱ+1)[μΔu+(μ+λ)divu]}hL25uhL23{ϱ(ϱ+1)[μΔu+(μ+λ)divu]}L25uhL2(ϱL5uL2+2uL33ϱL6)5uhL2(ϱL2122ϱL2125uL2+2uL2123uL2124ϱL2)5uhL2(1+t)-1545uhL2+(1+t)-1745uhL2(1+t)-112+(1+t)-25uhL22.

(2.26) |I5|3{[P(ϱ+1)ϱ+1-P(1)]ϱ}hL25uhL23{[P(ϱ+1)ϱ+1-P(1)]ϱ}L25uhL2(ϱL4ϱL2+ϱL33ϱL6)5uhL2ϱL2122ϱL2124ϱL25uhL2(1+t)-1545uhL2(1+t)-112+(1+t)-25uhL22.

For term I6, we can rewrite it as follows

(2.27) I6=-4(ϑ24ϱϱ+1Δϱ)h,4uh=-4(ϑ24ϱϱ+1Δϱ)+4(ϑ24ϱϱ+1Δϱ)l,4uh=-4(ϑ24ϱϱ+1Δϱh)-4(ϑ24ϱϱ+1Δϱl)+4(ϑ24ϱϱ+1Δϱ)l,4uh:=I6,1+I6,2+I6,3.

Similar to the proof of (2.19)(2.21), for I6,1, I6,2 and I6,3, it holds that

|I6,1|3(ϑ24ϱϱ+1Δϱh)L25uhL2(ϱL6ϱhL2+3ϱhL63ϱL3)5uhL2(ϱL2122ϱL2126ϱhL2+3ϱL2124ϱL2124ϱL2)5uhL2(1+t)-326ϱhL25uhL2+(1+t)-925uhL2(1+t)-112+(1+t)-32(6ϱhL22+5uhL22).

|I6,2|3(ϑ24ϱϱ+1Δϱl)L25uhL2(ϱL6ϱlL2+3ϱlL63ϱL3)5uhL2(ϱL5ϱL2+3ϱlL63ϱL3)5uhL2(ϱL2122ϱL2125ϱL2+3ϱL2124ϱL2124ϱL2)5uhL2(1+t)-1545uhL2+(1+t)-925uhL2(1+t)-112+(1+t)-25uhL22.

Using the properties of low-frequency decomposition, it holds that

|I6,3|3(ϑ24ϱϱ+1Δϱ)lL25uhL22(ϑ24ϱϱ+1Δϱ)L25uhL2(ϱL5ϱL2+3ϱL62ϱL3)5uhL2(ϱL2122ϱL2125ϱL2+2ϱL2123ϱL2124ϱL2)5uhL2(1+t)-1545uhL2(1+t)-112+(1+t)-25uhL22.

Putting the above estimates into (2.27) yields directly

(2.28) |I6|(1+t)-112+(1+t)-32(6ϱhL22+5uhL22).

For the terms I7, I8, we have

(2.29) |I7|3[ϑ24(|ϱ|2ϱ(1+ϱ)3-ϱΔϱ(1+ϱ)2-ϱ2ϱ(1+ϱ)2)]hL25uhL23[ϑ24(|ϱ|2ϱ(1+ϱ)3-ϱΔϱ(1+ϱ)2-ϱ2ϱ(1+ϱ)2)]L25uhL2(ϱL5ϱL2+4ϱL62ϱL3)5uhL22ϱL2123ϱL2125ϱL25uhL2(1+t)-1745uhL2(1+t)-112+(1+t)-35uhL22,

(2.30) |I8|3[11+ϱ((×B)×B)]hL25uhL23[11+ϱ((×B)×B)]L25uhL2(BL4BL2+BL33BL6)5uhL2BL2122BL2124BL25uhL2(1+t)-1545uhL2(1+t)-112+(1+t)-25uhL22.

Summing up (2.23)(2.26), (2.28)(2.30), we obtain

(2.31) |4G2h,4uh|(1+t)-112+(1+t)-32(5uhL22+6ϱhL22),

and we have

(2.32) |4G3h,4Bh|3(×(u×B))hL25BhL23(×(u×B))L25BhL2(uL4BL2+BL4uL2)5BhL2(uL2122uL2124BL2+BL2122BL2124uL2)5BhL2(1+t)-1545BhL2(1+t)-112+(1+t)-25BhL22.

For the term 〈 5G1h , ∇5 ϱh〉, we have

(2.33) ϑ245G1h,5ϱh=-ϑ245(ϱdivu+uϱ)h,5ϱh=-ϑ245(ϱdivu)h,5ϱh+ϑ24-5(uϱ)h,5ϱh:=I9+I10.

For the term I9, by using the properties of low-frequency and high-frequency decomposition, we get

(2.34) I9=ϑ24-5(ϱdivu)h,5ϱh=ϑ24-5(ϱdivu)+5(ϱdivu)l,5ϱh=ϑ24-5(ϱdivuh)-5(ϱdivul)+5(ϱdivu)l,5ϱh=I9,1+I9,2+I9,3.

For the term I9,1, by using integration by parts, we obtain

(2.35) |I9,1|4(ϱdivuh)L26ϱhL2(ϱL5uhL2+uhL34ϱL6)6ϱhL2(ϱL2122ϱL2125uhL2+uL2122uL2125ϱL2)6ϱhL2(1+t)-325uhL26ϱhL2+(1+t)-1546ϱhL2(1+t)-112+(1+t)-32(6ϱhL22+5uhL22).

For the term I9,2, we have

(2.36) |I9,2|4(ϱdivul)L26ϱhL2(ϱL5ulL2+ulL34ϱL6)6ϱhL2(ϱL2122ϱL2124uL2+uL2122uL2125ϱL2)6ϱhL2(1+t)-1546ϱhL2(1+t)-112+(1+t)-26ϱhL22.

For the term I9,3, we get

(2.37) |I9,3|4(ϱdivu)lL26ϱhL23(ϱdivu)L26ϱhL2(ϱL4uL2+uL33ϱL6)6ϱhL2(ϱL2122ϱL2124uL2+uL2122uL2124ϱL2)6ϱhL2(1+t)-1546ϱhL2(1+t)-112+(1+t)-26ϱhL22.

Putting estimates (2.35)(2.37) into (2.34), it arrives at

(2.38) |I9|(1+t)-112+(1+t)-32(6ϱhL22+5uhL22).

Similarly, for the term I10, we can rewrite it as follows

(2.39) I10=ϑ24-5(uϱ)h,5ϱh=ϑ24-5(uϱ)+5(uϱ)l,5ϱh=ϑ24-5(uϱh)-5(uϱl)+5(uϱ)l,5ϱh=I10,1+I10,2+I10,3.

Similar to the proof of (2.35), we have

(2.40) |I10,1|4(uϱh)L26ϱhL2(uL5ϱhL2+ϱhL4uL2)6ϱhL2(uL2122uL2125ϱhL2+2ϱL2123ϱL2124uL2)6ϱhL2(1+t)-1746ϱhL2+(1+t)-1546ϱhL2(1+t)-112+(1+t)-26ϱhL22,

(2.41) |I10,2|4(uϱl)L26ϱhL2(uL5ϱlL2+ϱlL4uL2)6ϱhL2(uL2122uL2124ϱL2+2ϱL2123ϱL2124uL2)6ϱhL2(1+t)-1546ϱhL2(1+t)-112+(1+t)-26ϱhL22,

and

(2.42) |I10,3|4(uϱ)lL26ϱhL23(uϱ)L26ϱhL2(uL4ϱL2+ϱL33uL6)6ϱhL2(uL2122uL2124ϱL2+ϱL2122ϱL2124uL2)6ϱhL2(1+t)-1546ϱhL2(1+t)-112+(1+t)-26ϱhL22.

It follows from the estimates (2.40)(2.42) and (2.39) that

(2.43) |I10|(1+t)-112+(1+t)-26ϱhL22.

Combining with (2.12), (2.22), (2.31)(2.32), (2.38) and (2.43), it arrives at

(2.44) 12ddt3|4ϱh|2+|4uh|2+|4Bh|2+ϑ24|5ϱh|2dx+(2μ+λ)3|5uh|2dx+ν3|5Bh|2dx(1+t)-112+(1+t)-2(4ϱhH12+5BhL22)+(1+t)-32(5uhL22+6ϱhL22).

In order to close the estimate, we will establish the dissipation estimate for ∇4 ϱh. Applying the operator 𝔉−1(1 − ϕ(ξ )) to ∇4(2.1)2, multiplying the resulting equality by ∇5 ϱh, integrating over ℝ3, we can deduce that

(2.45) ddt34uh5ϱhdx+5ϱhL22+ϑ246ϱhL22=5uhL22-μ5uh,6ϱh-(μ+λ)4divuh,6ϱh-4div(ϱu)h,6uh+4G2h,5ϱh.

Using the Young inequality, we can obtain

(2.46) |μ5uh,6ϱh|μ25uhL22+146ϱhL22.

Similarly, we have

(2.47) |(μ+λ)4divuh,6ϱh|(μ+λ)24divuhL22+146ϱhL22.

For 〈∇4div(ϱu)h, ∇5 ϱh〉, we obtain

(2.48) |4div(ϱu)h,5uh|(1+t)-112+(1+t)-325uhL22.

We notice that

(2.49) 4G2h,5ϱh=-4(uu)h,5ϱh+4{ϱ(ϱ+1)[μΔu+(μ+λ)divu]}h,5ϱh+-4{[P(ϱ+1)ϱ+1-P(1)]ϱ}h,5ϱh+-4(ϑ24ϱϱ+1Δϱ)h,5ϱh+4[ϑ24(|ϱ|2ϱ(1+ϱ)3-ϱΔϱ(1+ϱ)2-ϱ2ϱ(1+ϱ)2)]h,5ϱh+4[11+ϱ((×B)×B)]h,5ϱh:=j=16Jj.

The right-hand side of the above equation can be estimated as follows:

(2.50) |J1|3(uu)hL26ϱhL23(uu)L26ϱhL2(uL4uL2+3uL6uL3)6ϱhL2uL2122uL2124uL26ϱhL2(1+t)-1546ϱhL2(1+t)-112+(1+t)-26ϱhL22.

Similar to the proof of (2.38), we have

(2.51) |J2|(1+t)-112+(1+t)-32(6ϱhL22+5uhL22).

For J3, we get

(2.52) |J3|3{[P(ϱ+1)ϱ+1-P(1)]ϱ}hL26ϱhL23{[P(ϱ+1)ϱ+1-P(1)]ϱ}L26ϱhL2(ϱL4ϱL2+ϱL33ϱL6)6ϱhL2ϱL2122ϱL2124ϱL26ϱhL2(1+t)-1546ϱhL2(1+t)-112+(1+t)-26ϱhL22.

For J4, we can rewrite it as follows

(2.53) J4=ϑ243(ϱϱ+1Δϱ)h,6ϱh=ϑ243(ϱϱ+1Δϱ)-3(ϱϱ+1Δϱ)l,6ϱh=ϑ243(ϱϱ+1Δϱh)+3(ϱϱ+1Δϱl)-3(ϱϱ+1Δϱ)l,6ϱh=J4,1+J4,2+J4,3.

For J4,1, we obtain

(2.54) |J4,1|3(ϱϱ+1Δϱh)L26ϱhL2(ϱL6ϱhL2+3ϱhL63ϱL3)6ϱhL2(ϱL2122ϱL2126ϱhL2+3ϱL2124ϱL2124ϱhL2)6ϱhL2(1+t)-326ϱhL2+(1+t)-926ϱhL2(1+t)-112+(1+t)-326ϱhL22.

For J4,2, we get

(2.55) |J4,2|3(ϱϱ+1Δϱl)L26ϱhL2(ϱL6ϱlL2+3ϱL63ϱlL3)6ϱhL2(ϱL2122ϱL2125ϱL2+3ϱlL2124ϱlL2124ϱL2)6ϱhL2+(1+t)-1546ϱhL2(1+t)-112+(1+t)-26ϱhL22.

For J4,3, we get

(2.56) |J4,3|3(ϱϱ+1Δϱ)lL26ϱhL22(ϱϱ+1Δϱ)L26ϱhL2(ϱL5ϱL2+3ϱL62ϱL3)6ϱhL2(ϱL2122ϱL2125ϱL2+2ϱL2123ϱL2124ϱL2)6ϱhL2(1+t)-1546ϱhL2+(1+t)-1746ϱhL2(1+t)-112+(1+t)-26ϱhL22.

The combination of (2.53)(2.56) give rises to

(2.57) |J4|(1+t)-112+(1+t)-326ϱhL22,

(2.58) |J5|3[ϑ24(|ϱ|2ϱ(1+ϱ)3-ϱΔϱ(1+ϱ)2-ϱ2ϱ(1+ϱ)2)]hL26ϱhL23[ϑ24(|ϱ|2ϱ(1+ϱ)3-ϱΔϱ(1+ϱ)2-ϱ2ϱ(1+ϱ)2)]L26ϱhL2(ϱL5ϱL2+4ϱL62ϱL3)6ϱhL22ϱL2123ϱL2125ϱL26ϱhL2(1+t)-1746ϱhL2(1+t)-112+(1+t)-36ϱhL22,

and

(2.59) |J6|3[11+ϱ((×B)×B)]hL26ϱhL23[11+ϱ((×B)×B)]L26ϱhL2(BL4BL2+BL33BL6)6ϱhL2BL2122BL2124BL26ϱhL2(1+t)-1546ϱhL2(1+t)-112+(1+t)-26ϱhL22.

It follows from the estimates (2.50)(2.52) and (2.57)(2.59) that

(2.60) |3G2h,4ϱh|(1+t)-112+(1+t)-32(6ϱhL22+5uhL22),

which together with (2.46)(2.48) implies immediately

(2.61) ddt34uh5ϱhdx+125ϱhL22+ϑ246ϱhL22(1+t)-112+5uhL22+(1+t)-32(6ϱhL22+5uhL22).

Last, we will close the estimate to prove the decay rate (1.8). Choosing sufficient large time T1 and positive constant D0, we define the temporary energy functional

(2.62) (t)=D0(4(ϱ,u,B)hL22+ϑ225ϱhL22)+34uh5ϱhdx,

for t ≥ 0, where it is clear that ℰ(t) is equivalent to 4ϱhH12+4(u,B)hL22 since D0 is large enough.

Combining the estimates (2.44) with (2.61) and using Lemma 2.3, for tT1, one may deduce that

(2.63) ddt(t)+4ϱh(t)H12+4uh(t)L22+4Bh(t)L22(1+t)-112,

where we have used the fact that T1 is large enough. On the other hand, it is easy to see that

(2.64) 4ϱh(t)H12+4uh(t)L22+4Bh(t)L22C(t).

Therefore, (1.8) follows from (2.63), (2.64), Lemma 2.3 and Gronwall's argument immediately.

3 Proof of Theorem 1.3

This section is concerned with the optimal convergence rate of the solution, which is stated on Theorem 1.3. First, we will derive optimal convergence rate on ‖∇5 ϱlL2.

Step 1. Low–frequency L2 energy estimate. By virtue of the equation (2.1)1, Theorem 1.2, Duhamel’s principle, Plancherel theorem, Hölder’s inequality and Hausdorff–Young’s inequality, we have

(3.1) 5ϱl(t)L2 (1+t)-134(ϱ,u,B)(0)L1+0t2(1+t-τ)-134G(τ)L1dτ+t2t(1+t-τ)-54|ξ|4G^l(τ)Ldτ.

For ‖G(τ)‖L1, we can see the proof of (2.8). On the other hand, we can use (1.8) to bound the third term on the right-hand side of (3.1) as follows

(3.2) |ξ|4G^l(τ)L4Gl(τ)L13(div(ϱu))(τ)L1+3(uu)(τ)L1+2(ϱϱ+1[μΔu+(μ+λ)divu])(τ)L1+3([P(ϱ+1)ϱ+1-P(1)]ϱ)(τ)L1+2(ϑ24ϱϱ+1Δϱ)(τ)L1+3(ϑ24(|ϱ|2ϱ(1+ϱ)3-ϱΔϱ(1+ϱ)2-ϱ2ϱ(1+ϱ)2))(τ)L1+3(11+ϱ((×B)×B))(τ)L1+3(×(u×B)(τ))L1(ϱ,u)(τ)L24(ϱ,u)(τ)L2+2u(τ)L22+ϱ(τ)L25ϱ(τ)L2+ϱ(τ)L25ϱ(τ)L2+ϱ(τ)L24ϱ(τ)L2+2ϱ(τ)L23ϱ(τ)L2+B(τ)L23B(τ)L2+(u,B)(τ)L24(u,B)(τ)L2(1+t)-72.

Substituting (3.2) into (3.1), it arrives at

(3.3) 5ϱlL2(1+t)-134.

Step 2. High–frequency L2 energy estimate. In this part, we will make full use of (1.8) to achieve this goal. It is not difficult to find that the following calculations are consistent with the second section except replacing the L2 decay rate (1+t)-94 by (1+t)-114 for the fourth order spatial derivative of solution. Similar to the proof of (2.48), we have

(3.4) ddt(t)+4ϱh(t)H12+4uh(t)L22+4Bh(t)L22(1+t)-132.

Finally, we conclude that

(3.5) 4ϱh(t)H1+4uh(t)L2+4Bh(t)L2(1+t)-134,

which together with (3.3) implies (1.9) directly.

4 Analytic tools

Now, we state some auxiliary lemmas, which will be frequently used in this paper.

Lemma 4.1

(Gagliardo-Nirenberg’s inequality) Given 2 ≤ p ≤ + ∞ and 0 ≤ k, mℓ, then for any fH(ℝ3), we have

kfLpmfL2α𝓁fL21-α,

where α ∈ [0, 1] satisfies

k3-1p=(m3-12)α+(𝓁3-12)(1-α).

Proof

This is the special case of [23].

Lemma 4.2

Let k ≥ 1 be an integer, then it holds that

k(fg)LpfLp1kgLp2+gLp3kfLp4,

where p, p2, p3 ∈ (1, +∞) and

1p=1p1+1p2=1p3+1p4.

Proof

For p = p2 = p3 = 2, it can be proved by using Lemma 4.1. For the general case, one may refer to [12]

Lemma 4.3

If fLr (ℝ3) for any 2 ≤ r ≤ ∞, then we have

flLr+fhLrfLr.

Proof

For 2 ≤ r ≤ ∞, by Young inequality’s for convolutions, for the low frequency, it holds

flLr𝔉-1ϕL1fLrfLr,

and hence

fhLrfLr+flLrfLr.

Lemma 4.4

Let r1, r2 > 0, then it holds that

0t(1+t-s)-r1(1+s)-r2dsC(1+t)-min{r1,r2,r1+r2-1-},

for any arbitrarily small ε > 0.

Acknowledgments

This work is partially supported by Guangxi Natural Science Foundation #2019JJG110003, #2019AC20214, and National Natural Science Foundation of China #11771150.

  1. Conflict of interest: Authors state no conflict of interest.

References

[1] B. Karine and Z. Enrique, Large time asymptotics for partially dissipative hyperbolic systems, Arch. Ration. Mech. Anal. 199 (2011), no.1, 177–227.10.1007/s00205-010-0321-ySearch in Google Scholar

[2] Q. Chen and Z. Tan, Global existence and convergence rates of smooth solutions for the compressible magnetohydrody namical equations, Nonlinear Anal. 72 (2010), no.12, 4438–4451.10.1016/j.na.2010.02.019Search in Google Scholar

[3] R. J. Duan, S. Ukai, Y. Yang and H. J. Zhao, Optimal convergence rates for the compressible Navier-Stokes equations with potential forces, Math. Models Methods Appl. Sci. 17 (2007), no.5, 737–758.10.1142/S021820250700208XSearch in Google Scholar

[4] R. J. Duan, H. X. Liu, S. Ukai and T. Yang, Optimal LpLq convergence rates for the compressible Navier-Stokes equations with potential force, J. Differential Equations. 238 (2007), no.1, 220–233.10.1016/j.jde.2007.03.008Search in Google Scholar

[5] R. J. Duan, Green’s function and large time behavior of the Navier-Stokes-Maxwell system, Anal. Appl. 10 (2012), no.2, 133–197.10.1142/S0219530512500078Search in Google Scholar

[6] D. K. Ferry and J. R. Zhou, Form of the quantum potential for use in hydrodynamic equations for semiconductor device modeling, Phys Rev B Condens Matter. 48 (1993), no.11, 7944–7950.10.1103/PhysRevB.48.7944Search in Google Scholar

[7] J. C. Gao, Z. Y. Luy and Z. A. Yao, Lower bound of decay rate for higher order derivatives of solution to the compressible quantum magnetohydrodynamic model, Math. Methods Appl. Sci, 44 (2021), no.5, 3686–3704.10.1002/mma.6974Search in Google Scholar

[8] J. Grant, Pressure and stress tensor expressions in the fluid mechanical formulation of the Bose condensate equations, J. Phys. A. 6 (1973), no.11, L151–L153.10.1088/0305-4470/6/11/001Search in Google Scholar

[9] Y. Guo and Y. J. Wang, Decay of dissipative equations and Negative Sobolev spaces, Comm. Partial Differential Equations. 37 (2012), no.12, 2165–2208.10.1080/03605302.2012.696296Search in Google Scholar

[10] F. Haas, A magnetohydrodynamic model for quantum plasmas, Phys Plasmas. 12 (2005), no.6, 062117–062125.10.1063/1.1939947Search in Google Scholar

[11] D. Hoff and K. Zumbrun, Pointwise decay estimates for multidimensional Navier-Stokes diffusion waves, Z. Angew. Math. Phys. 48 (1997), no.4, 597–614.10.1007/s000330050049Search in Google Scholar

[12] N. Ju, Existence and uniqueness of the solution to the dissipative 2D quasi-geostrophic equations in the Sobolev space, Comm. Math. Phys. 251 (2004), no.2, 365–376.10.1007/s00220-004-1062-2Search in Google Scholar

[13] Y. Kagei and T. Kobayashi, On large time behavior of solutions to the compressible Navier-Stokes equations in the half space in ℝ3, Arch. Ration. Mech. Anal. 165 (2002), no.2, 89–159.10.1007/s00205-002-0221-xSearch in Google Scholar

[14] Y. Kagei and T. Kobayashi, Asymptotic behavior of solutions of the compressible Navier-Stokes equations on the half space, Arch. Rational Mech. Anal. 177 (2005), no.2, 231–330.10.4064/bc70-0-8Search in Google Scholar

[15] T. Kobayashi, Some estimates of solutions for the equations of motion of compressible viscous fluid in the three–dimensional exterior domain, J. Differential Equations. 184 (2002), no.2, 587–619.10.1006/jdeq.2002.4158Search in Google Scholar

[16] T. Kobayashi and Y. Shibata, Decay estimates of solutions for the equations of motion of compressible viscous and heat-conductive gases in an exterior domain in ℝ3, Comm. Math. Phys. 200 (1999), no.3, 621–659.10.1007/s002200050543Search in Google Scholar

[17] T. Kobayashi and Y. Shibata, Remark on the rate of decay of solutions to linearized compressible Navier-Stokes equations, Pacific J. Math. 207 (2002), no.1, 199–234.10.2140/pjm.2002.207.199Search in Google Scholar

[18] H. L. Li, A. Matsumura and G. J. Zhang, Optimal decay rate of the compressible Navier-Stokes-Poisson system in ℝ3, Arch. Ration. Mech. Anal. 196 (2010), no.2, 681–713.10.1007/s00205-009-0255-4Search in Google Scholar

[19] T. P. Liu and W. K. Wang, The pointwise estimates of diffusion waves for the Navier-Stokes equations in odd multidimensions, Comm. Math. Phys. 196, (1998), 145–173.10.1007/s002200050418Search in Google Scholar

[20] Q. Q. Liu and C. J. Zhu, Asymptotic stability of stationary solutions to the compressible Euler Maxwell equations, Indiana Univ. Math. J. 62 (2013), no.4, 1203–1235.10.1512/iumj.2013.62.5047Search in Google Scholar

[21] A. Matsumura and T. Nishida, The initial value problem for the equation of motion of compressible viscous and heat–conductive fluids, Proc. Japan Acad. Ser. A Math. Sci. 55 (1979), no.9, 337–342.10.3792/pjaa.55.337Search in Google Scholar

[22] A. Matsumura and T. Nishida, Initial boundary value problems for the equations of motion of compressible viscous and heat-conductive fluids, Comm. Math. Phys. 89 (1983), no.4, 445–464.10.1007/BF01214738Search in Google Scholar

[23] L. Nirenberg, On elliptic partial differential equations, Ann. Sc. Norm. Super. Pisa Cl. Sci. 13 (1959), no.2, 115–162.10.1007/978-3-642-10926-3_1Search in Google Scholar

[24] G. Ponce, Global existence of small solution to a class of nonlinear evolution equations, Nonlinear Anal. 9 (1985), no.5, 399–418.10.1016/0362-546X(85)90001-XSearch in Google Scholar

[25] X. K. Pu and B. L. Guo, Global existence and convergence rates of smooth solutions for the full compressible MHD equations, Z. Angew. Math. Phys. 64 (2013), no.3, 519–538.10.1007/s00033-012-0245-5Search in Google Scholar

[26] X. K. Pu and B. L. Guo, Optimal decay rate of the compressible quantum Navier-Stokes equations, Ann. Appl. Math. 32 (2016), no.3, 275–287.Search in Google Scholar

[27] X. K. Pu and X. L. Xu, Decay rates of the magnetohydrodynamic model for quantum plasmas, Z. Angew. Math. Phys. 68 (2017), no.1.10.1007/s00033-016-0762-8Search in Google Scholar

[28] X. K. Pu and X. L. Xu, Asymptotic behaviors of the full quantum hydrodynamic equations, J. Math. Anal. Appl. 454 (2017), no.1, 219–245.10.1016/j.jmaa.2017.04.053Search in Google Scholar

[29] Z. Tan, Y. J. Wang and Y. Wang, Stability of steady states of the Navier-Stokes-Poisson equations with non-flat doping profile, SIAM J. Math. Anal. 47 (2015), no.1, 179–209.10.1137/130950069Search in Google Scholar

[30] S. Ukai, T. Yang and H. J. Zhao, Convergence rate for the compressible Navier-Stokes equations with external force, J. Hyperbolic Differ. Equ. 3 (2006), no.3, 561–574.10.1142/S0219891606000902Search in Google Scholar

[31] Y. J. Wang and Z. Tan, Optimal decay rates for the compressible fluid models of Korteweg type, J. Math. Anal. Appl. 379 (2011), no.1, 256–271.10.1016/j.jmaa.2011.01.006Search in Google Scholar

[32] J. Wang, C. G. Xiao and Y. H. Zhang, Optimal large time behavior of the compressible Navier–Stokes–Korteweg system in ℝ3, Appl. Math. Lett. 120, (2021), 107274.10.1016/j.aml.2021.107274Search in Google Scholar

[33] G. C. Wu, Y. H. Zhang and W. Y. Zou, Optimal time-decay rates for the 3D compressible Magnetohydrodynamic flows with discontinuous initial data and large oscillations, J. Lond. Math. Soc. 103 (2020), no.3, 817–845.10.1112/jlms.12393Search in Google Scholar

[34] R. Wyatt, Quantum Dynamics with Trajectories, Introduction to Quantum Hydrodynamics, With contributions by Corey J. Trahan, Interdisciplinary Applied Mathematics. 28 (2005).Search in Google Scholar

[35] X. Y. Xi, X. K. Pu and B. L. Guo, Long-time behavior of solutions for the compressible quantum magnetohydrodynamic model in ℝ3, Z. Angew. Math. Phys. 70 (2019), no.1, 7.10.1007/s00033-018-1049-zSearch in Google Scholar

[36] B. Q. Xie, X. Y. Xi and B. L. Guo, Long-time behavior of solutions for full compressible quantum model in ℝ3, Appl. Math. Lett. 80, (2018), 54–58.10.1016/j.aml.2018.01.008Search in Google Scholar

Received: 2021-09-24
Accepted: 2021-12-19
Published Online: 2022-02-09

© 2021 Juan Wang et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 9.6.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2021-0219/html
Scroll to top button