Processing math: 100%
Research article Topical Sections

In-situ reactions in hybrid aluminum alloy composites during incorporating silica sand in aluminum alloy melts

  • Received: 24 May 2016 Accepted: 18 July 2016 Published: 19 July 2016
  • In order to gain a better understanding of the reactions and strengthening behavior in cast aluminum alloy/silica composites synthesized by stir mixing, experiments were conducted to incorporate low cost foundry silica sand into aluminum composites with the use of Mg as a wetting agent. SEM and XRD results show the conversion of SiO2 to MgAl2O4 and some Al2O3 with an accompanying increase in matrix Si content. A three-stage reaction mechanism proposed to account for these changes indicates that properties can be controlled by controlling the base Alloy/SiO2/Mg chemistry and reaction times. Experimental data on changes of composite density with increasing reaction time and SiO2 content support the three-stage reaction model. The change in mechanical properties with composition and time is also described.

    Citation: Afsaneh Dorri Moghadam, J.B. Ferguson, Benjamin F. Schultz, Pradeep K. Rohatgi. In-situ reactions in hybrid aluminum alloy composites during incorporating silica sand in aluminum alloy melts[J]. AIMS Materials Science, 2016, 3(3): 954-964. doi: 10.3934/matersci.2016.3.954

    Related Papers:

    [1] Shuai Yuan, Sitong Chen, Xianhua Tang . Normalized solutions for Choquard equations with general nonlinearities. Electronic Research Archive, 2020, 28(1): 291-309. doi: 10.3934/era.2020017
    [2] Jun Wang, Yanni Zhu, Kun Wang . Existence and asymptotical behavior of the ground state solution for the Choquard equation on lattice graphs. Electronic Research Archive, 2023, 31(2): 812-839. doi: 10.3934/era.2023041
    [3] Xiaoguang Li . Normalized ground states for a doubly nonlinear Schrödinger equation on periodic metric graphs. Electronic Research Archive, 2024, 32(7): 4199-4217. doi: 10.3934/era.2024189
    [4] Weiyu Li, Hongyan Wang . Dynamics of a three-molecule autocatalytic Schnakenberg model with cross-diffusion: Turing patterns of spatially homogeneous Hopf bifurcating periodic solutions. Electronic Research Archive, 2023, 31(7): 4139-4154. doi: 10.3934/era.2023211
    [5] Haijun Luo, Zhitao Zhang . Existence and stability of normalized solutions to the mixed dispersion nonlinear Schrödinger equations. Electronic Research Archive, 2022, 30(8): 2871-2898. doi: 10.3934/era.2022146
    [6] Xiaoyong Qian, Jun Wang, Maochun Zhu . Existence of solutions for a coupled Schrödinger equations with critical exponent. Electronic Research Archive, 2022, 30(7): 2730-2747. doi: 10.3934/era.2022140
    [7] Jingjing Zhang, Ting Zhang . Local well-posedness of perturbed Navier-Stokes system around Landau solutions. Electronic Research Archive, 2021, 29(4): 2719-2739. doi: 10.3934/era.2021010
    [8] Ping Yang, Xingyong Zhang . Existence of nontrivial solutions for a poly-Laplacian system involving concave-convex nonlinearities on locally finite graphs. Electronic Research Archive, 2023, 31(12): 7473-7495. doi: 10.3934/era.2023377
    [9] Lingzheng Kong, Haibo Chen . Normalized solutions for nonlinear Kirchhoff type equations in high dimensions. Electronic Research Archive, 2022, 30(4): 1282-1295. doi: 10.3934/era.2022067
    [10] Quanqing Li, Zhipeng Yang . Existence of normalized solutions for a Sobolev supercritical Schrödinger equation. Electronic Research Archive, 2024, 32(12): 6761-6771. doi: 10.3934/era.2024316
  • In order to gain a better understanding of the reactions and strengthening behavior in cast aluminum alloy/silica composites synthesized by stir mixing, experiments were conducted to incorporate low cost foundry silica sand into aluminum composites with the use of Mg as a wetting agent. SEM and XRD results show the conversion of SiO2 to MgAl2O4 and some Al2O3 with an accompanying increase in matrix Si content. A three-stage reaction mechanism proposed to account for these changes indicates that properties can be controlled by controlling the base Alloy/SiO2/Mg chemistry and reaction times. Experimental data on changes of composite density with increasing reaction time and SiO2 content support the three-stage reaction model. The change in mechanical properties with composition and time is also described.


    In this paper, we are looking for solutions to the following Choquard equation

    {Δu=λu+(IαF(u))f(u)+μ|u|q2uinRN,RN|u|2dx=a>0, (1.1)

    where N3, α(0,N), F(s):=s0f(t)dt, μ>0, 2<q¯q:=2+4N, a>0 is a given mass, and λR appears as an unknown Lagrange multiplier. Iα:RN{0}R is the Riesz potential defined by

    Iα(x):=Aα|x|NαwithAα=Γ(Nα2)2απN2Γ(α2).

    The Choquard equation

    Δu+u=(IαF(u))f(u)inRN, (1.2)

    has been studied extensively for its profound physical backgrounds. In particular, when N=3, α=2, and F(s)=s22, Eq (1.2) turns into

    Δu+u=(I2|u|2)uinR3, (1.3)

    which was introduced by Pekar [1] to describe the quantum theory of a polaron at rest and then used by Choquard [2] to study steady states of the one one-component plasma approximation to the Hartree-Fock theory. Also, Eq (1.3) reemerged as a model of self-gravitating matter [3,4], and in that context it is known as the Schrödinger-Newton equation.

    The pioneering mathematical research dates back to Lieb [2], in which the author proved the existence and uniqueness for Eq (1.3) by variational methods. Later, Lions [5] obtained the existence of normalized solutions. In the homogeneous nonlinearity case of Eq (1.2) with F(s)=1p|s|p, Moroz and Van Schaftingen in [6] established the existence of ground states to Eq (1.2) with an optimal range N+αN<p<N+αN2. Moreover, Ghimenti and Van Schaftingen in [7] obtained solutions which are odd with respect to a hyperplane of RN. The existence of saddle type nodal solution for the Choquard equation was proven in the study [8]. For a more general nonlinearity, Moroz and Van Schaftingen [9] proved Eq (1.2) has a ground solution when the nonlinearity satisfies Berestycki-Lions type condition and N3, see also [10] for the case N=2. For the Choquard equation with a local nonlinear perturbation

    Δu+u=(Iα|u|p)|u|p2u+|u|q2u,inRN. (1.4)

    When N=3, 0<α<1, p=2, and 4q<6, Chen and Gao in [11] obtained the existence of solutions of Eq (1.4). For q=2 in Eq (1.4), Seok in [12] constructed a family of nontrivial solutions. In [13], the authors studied Eq (1.4) with a general local nonlinearity f(x,u) subcritical type instead of |u|q2u.

    From a physical point of view, it is interesting to find solutions with prescribed L2 norm, since there is a conservation of mass. Solutions of this type are often referred to as normalized solutions. In recent years, normalized solutions to nonlinear elliptic problems have attracted much attention from researchers. In [14], Jeanjean using a mountain pass structure for a stretched functional to consider the equation

    Δu=λu+f(u),inRN. (1.5)

    He proved the existence of at least one normalized solution of Eq (1.5) in a purely L2 supercritical case. Later, Bartsch and Valeriola [15] obtained the existence of infinitely many normalized solutions by a new linking geometry for the stretched functional. For more general nonlinearity f(s) has L2 subcritical {growth, Shibata [16] obtained the existence and nonexistence of normalized ground state solution of} Eq (1.5) via minimizing method, Jeanjean and Lu in [17] showed the existence of nonradial normalized solutions for any N4. For more results on normalized solutions for Schrödinger equations by variational methods, we would like to refer [18,19,20,21,22,23]. Also, for the evolution equations with the singular potentials whose the steady state equations are the nonlinear elliptic equations, see [24].

    Concerning the normalized solutions of Choquard equations, Li and Ye in [25] considered the existence of normalized solutions to the following equation

    Δu=λu+(IαF(u))f(u),xRN, (1.6)

    under a set of assumptions on f, where f takes the special case f(s)=C1|s|r2s+C2|s|p2s requires that N+2+αN<rp<N+αN2. Later, Yuan et al. [26] generalized the results in [25] to more general fC(R,R). In [27], Bartsch et al. obtained the existence of infinitely many normalized solutions of (1.6). For F(u)=|u|pp, Ye in [28] studied the qualitative properties including existence and nonexistence of minimizers of the functional associated to the Eq (1.6). Yao et al. in [29] considered normalized solutions for the following problems

    Δu+λu=γ(Iα|u|p)|u|p2u+μ|u|q2u,inRN. (1.7)

    Under different assumptions on γ,μ,p, and q, they proved several existence, multiplicity, and nonexistence results. For more related topics, we refer the reader to [30,31,32,33].

    Motivated by the above papers, it is natural to ask if the nonlinearity f(u) in Eq (1.1) satisfies general growth assumptions, and if the normalized ground states still exist. In the present paper, we attempt to study this kind of problem. In order to prove the existence of normalized ground state solutions to Eq (1.1), assuming that nonlinearity fC(R,R) satisfies the growth assumptions:

    (f1) lims0f(s)|s|αN=0,

    (f2) lim|s|+f(s)|s|2+αN=0,

    (f3) f is odd and f does not change sign on (0,+),

    (f4) lims0F(s)|s|N+2+αN=+,

    (f5) lim sups0F(s)|s|N+2+αN<+.

    To find solutions of Eq (1.1), we define functional Iμ(u):H1(RN)R by

    Iμ(u)=12RN|u|2dx12RN(IαF(u))F(u)dxμqRN|u|qdx.

    For a>0, set

    S(a)={uH1(RN):|u|22=a}.

    Since fC(R,R) satisfies (f1) and (f2), using the Hardy-Littlewood-Sobolev inequality, we see that IμC1(H1(RN),R). Normalized solutions of Eq (1.1) can be obtained by looking for critical points of Iμ on the constraint S(a). It is standard that each critical point uS(a) of Iμ|S(a), corresponds a Lagrangian multiplier λR such that (u,λ) solves Eq (1.1). We will be interested in ground state solutions, and following [34], we say that uS(a) is a normalized ground state solution to Eq (1.1) if (Iμ|S(a))(u)=0 and

    Iμ(u)=inf{Iμ(v):vS(a),(Iμ|S(a))(v)=0}.

    In particular, if uS(a) is a minimizer of the minimization problem

    m(a)=infuS(a)Iμ(u),

    then u is a critical point of Iμ|S(a) as well as a normalized ground state to Eq (1.1).

    Our main results dealing with problem (1.1) are the following:

    Theorem 1.1. Assume that (f1)–(f3) hold. Let μ>0 and 2<q<¯q. Then, for any a>0, there exists a global minimizer ˜u with respect to m(a), which solves Eq (1.1) for some ˜λ<0. Moreover, ˜u is a ground state solution of Eq (1.1) which has constant sign, is radially symmetric with respect to some point in RN, and is decreasing.

    Theorem 1.2. Assume that (f1)–(f3) hold and q=¯q. For any μ>0 small enough, there exist a a0=a0(μ)>0 and a[0,a0) such that:

    (i) for any a(0,a), there is no global minimizer with respect to m(a).

    (ii) for any a(a,a0), there exists a global minimizer u with respect to m(a), which solves Eq (1.1) for some λ<0. Moreover, u is a ground state solution of Eq (1.1) which has constant sign, is radially symmetric with respect to some point in RN, and is decreasing.

    (iii) if (f4) holds, then a=0, and if (f5) holds, then a>0.

    Remark 1. The value a0=a0(μ)>0 is explicit and is given in Lemma 4.1. In particular, a0>0 can be taken arbitrary large by taking μ>0 small enough.

    The following result positively answers the existence of global minimizer with respect to m(a) for the sharp threshold a>0.

    Theorem 1.3. Assume that (f1)(f3) and

    (f6) lim sups0F(s)|s|N+2+αN=0 holds.

    Let q=¯q. For a=a, there exists a global minimizer v with respect to m(a)=0, which solves Eq (1.1) for some λ<0. In particular, v is a ground state solution of Eq (1.1) which has constant sign and is radially symmetric with respect to some point in RN.

    To the best of our knowledge, the main results in this paper are new. This is a complement of the results for Choquard equations about the existence of normalized solutions. Our main theorems can be viewed as an extension of some results in [25,29] to more general cases. In our setting, we only consider the situation when the constraint functional Iμ|S(a) is bounded from below and is coercive. As we will see, the existence of normalized states of Eq (1.1) are strongly effected by further assumptions on the exponent q. We are first devoted to prove the existence of ground state solutions of Eq (1.1) with 2<q<2+4N by application of the concentration-compactness principle [35]. In this case, m(a)<0 for any a>0 and the strict subadditivity inequality

    m(a+b)<m(a)+m(b)foralla,b>0 (1.8)

    holds, which permits us to exclude the dichotomy of the minimizing sequence. However, in the case of q=¯q:=2+4N, compare to [29], for a general f, the strict subadditivity inequality (1.8) does not hold, and m(a)<0 for all a>0 may not be satisfied. This prevents us from using the concentration-compactness principle in a standard way. In the proof, we adopt some ideas in [16,19] to recover the compactness of minimizing sequence with respect to m(a).

    The paper is organized as follows. In Section 2, we introduce the variational framework and give some preliminary results. In Section 3, we discuss the case of 2<q<¯q and prove Theorem 1.1. In Section 4, we deal with the case of q=¯q and prove Theorems 1.2 and 1.3.

    In this paper, we will use the following notations:

    H1(RN) is the usual Sobolev space endowed with the norm u2=RN(|u|2+|u|2)dx.

    Ls(RN), for 1s<, denotes the Lebesgue space with the norm |u|ss=RN|u|sdx.

    For any r>0 and xRN, Br(x) denotes the ball of radius of r centered at the x.

    The letters C, C0,C,C,C1,C2 are positive (possibly different) constants.

    on(1) denotes the vanishing quantities as n.

    In this section, we give some results which will be useful in forthcoming sections. First, let us recall the following Hardy-Littlewood-Sobolev inequality which will be frequently used throughout the paper.

    Lemma 2.1. (See [[36], Theorem4.3]). Suppose α(0,N), and s,r>1 with 1r+1s=1+αN. Let fLs(RN) and gLr(RN). Then, there exists a constant C(N,α,s,r)>0 such that

    |RNRNf(x)|xy|αNg(y)dxdy|C(N,α,s,r)|f|s|g|r.

    In particular, if r=s=2NN+α, then

    C(N,α,s,r)=Cα:=πNα2Γ(α2)Γ1(N+α2)[Γ(N2)Γ1(N)]αN.

    Let N1 and N+αN<p<2α:=N+αN2. Then, we introduce the following Gagliardo-Nirenberg inequality of Hartree type ([28])

    RN(Iα|u|p)|u|pdx¯CN,p|u|NpNα2|u|N+αp(N2)2. (2.1)

    We also recall the following Gagliardo-Nirenberg inequality. For p(2,2) and uH1(RN),

    |u|pCN,p|u|γp2|u|1γp2,whereγp=N(p2)2p. (2.2)

    Lemma 2.2. Assume that (f1)-(f2) hold. Let {un}H1(RN) be a bounded sequence. If either limn|un|2=0 or limn|un|2(N+α+2)N+α=0 holds, then

    limnRN(IαF(un))F(un)dx=0.

    Proof. By (f1) and (f2), for any ε>0, there exists Cε>0 such that

    |F(s)|ε|s|N+αN+Cε|s|N+2+αN.

    Then, using Lemma 2.1, for uH1(RN) we obtain

    |RN(IαF(u))F(u)dx|ε2RN(Iα|u|N+αN)|u|N+αNdx+2εCεRN(Iα|u|N+αN)|u|N+α+2Ndx+C2εRN(Iα|u|N+2+αN)|u|N+α+2NdxC1ε2|u|2(N+2)N2+C2C2ε|u|2(N+2+α)N2(N+2+α)N+α. (2.3)

    By Eq (2.2), we have

    |u|2(N+2+α)N2(N+2+α)N+αC3|u|22|u|2(2+α)N2. (2.4)

    If limn|un|2=0, by Eqs (2.3), (2.4), and the boundedness of {un}, we get

    limnRN(IαF(un))F(un)dx=0.

    If limn|un|2(N+α+2)N+α=0, by Eq (2.3), we have

    lim supn|RN(IαF(u))F(u)dx|C1ε2|un|2(N+2)N2.

    Since {un} is bounded in H1(RN) and ε>0 is arbitrary, the conclusion holds. The proof is completed.

    In our subsequent arguments, we will use the following nonlocal version of the Brezis-Lieb lemma.

    Lemma 2.3. (See [[37], Lemma 2.2]). Assume α(0,N) and there exists a constant C>0 such that

    |f(s)|C(|s|αN+|s|2+αN2),sR.

    Let {un}H1(RN) be such that unu weakly in H1(RN) and almost everywhere in RN as n. Then,

    RN(IαF(un))F(un)dx=RN(IαF(unu))F(unu)dx+RN(IαF(u))F(u)dx+on(1).

    Using a similar argument as the proof of ([9] Theorem 3), we have the following Pohoˇzaev of Eq (1.1).

    Lemma 2.4. Assume that N3 and α(0,N). If fC(R,R) satisfies (f1) and (f2), and if (u,λ)H1(RN)×R solves problem (1.1), then

    P(u)=N22|u|22N2λ|u|22N+α2RN(IαF(u))F(u)dxNμq|u|qq=0.

    For each uS(a) and t>0, we define the scaling function

    ut(x):=tN2u(tx).

    It is clear that utS(a) and

    Iμ(ut)=t22|u|2212tN+αRN(IαF(tN2u))F(tN2u)dxμqtN(q2)2|u|qq.

    In this section, we deal with the case 2<q<¯q and prove Theorem 1.1.

    Lemma 3.1. Assume that (f1) and (f2) hold. For any 2<q<¯q and a,μ>0, we have

    <m(a)=infuS(a)Iμ(u)<0.

    Proof. By (f1) and (f2), for any ε>0, there exists Cε>0 such that

    |F(s)|Cε|s|N+αN+ε|s|N+2+αN.

    Using Lemma 2.1 and Eq (2.2), we obtain, for uH1(RN)

    |RN(IαF(u))F(u)dx|C3C2ε|u|2(N+2)N2+C4ε2|u|2(α+2)N2|u|22. (3.1)

    Then, for any uS(a), by Eqs (2.2) and (3.1), we get

    Iμ(u)12|u|22C3C2εaN+2NC4ε2aα+2N|u|22μCqN,qqaq(1γq)2|u|qγq2. (3.2)

    Choosing ε=(4C4a2+αN)12, it follows from Eq (3.2) that

    Iμ(u)14|u|22CaN+2NμCqN,qq1aq(1γq)2|u|qγq2

    for every uS(a). Since 2<q<¯q, we see that 0<qγq<2, and hence Iμ is coercive on S(a), which provides that m(a)>.

    On the other hand, for uS(a)

    Iμ(ut)t22|u|22μqtN(q2)2|u|qq=tN(q2)2(12t2N(q2)2|u|22μq|u|qq).

    Noticing that 2<q<¯q, we have 2N(q2)2>0, and hence Iμ(ut)<0 for every uS(a) with t>0 small enough. Therefore, we have that m(a)<0 for any a>0.

    Since m(a)<0 for any a>0, we can give the following strict sub-additivity.

    Lemma 3.2. Let a1,a2>0 be such that a1+a2=a. Then,

    m(a)<m(a1)+m(a2).

    Proof. For uS(a) and θ>1, we set ¯u(x)=u(θ1Nx). Then, ¯u(x)S(θa). Let {un}S(a) be a minimizing sequence for m(a). Since θ>1, we have

    m(θa)Iμ(¯un)=θ12N2|un|22θ1+αN2RN(IαF(un))F(un)dxμθq|un|qq<θIμ(un)=θm(a)+on(1).

    As a consequence,

    m(θa)θm(a),

    with equality if and only if

    limn(Iμ(¯un)θIμ(un))=0. (3.3)

    But this is can not occur. Otherwise, by Eq (3.3), we find

    limn(θ2N12|un|22+1θαN2RN(IαF(un))F(un)dx)=0.

    By Lemma 3.1, Iμ is coercive on S(a) and we have {un} is bounded in H1(RN). It follows from θ>1 that

    limn|un|22=0=limnRN(IαF(un))F(un)dx.

    Combining Eq (2.2) we get limn|un|qq=0. Then, by Lemma 3.1, we obtain

    0>m(a)=limnIμ(un)=0,

    a contradiction. Thus, we have the strict inequality

    m(θa)<θm(a)foranyθ>1. (3.4)

    Next, we show that m(a)<m(a1)+m(a2). We may assume that a1a2, by Eq (3.4) we have

    m(a)=m(aa1a1)<aa1m(a1)=m(a1)+a2a1m(a1)m(a1)+m(a2).

    The proof is completed.

    Lemma 3.3. Assume that (f1) and (f2) hold and a,μ>0. Let {un}H1(RN) be a sequence such that

    limnIμ(un)=m(a),limn|un|22=a. (3.5)

    Then, taking a subsequence if necessary, there exist ˜uS(a) and a family {yn}RN such that un(+yn)˜u in H1(RN) as n. Specifically, ˜u is a global minimizer.

    Proof. Since {un}H1(RN) satisfies Eq (3.5), it is easy to see that {un} is bounded in H1(RN). From the concentration-compactness lemma [35], there exists a subsequence of {un} (denoted in the same way) satisfying one of the three following possibilities:

    vanishing: for all R>0

    limnsupyRNBR(y)|un|2dx=0;

    dichotomy: there exists a constant b(0,a), sequences {u(1)n}, {u(2)n} bounded in H1(RN) such that as n

    {|un(u(1)n+u(2)n)|p0,for2p<2,|u(1)n|22b,|u(2)n|22ab,dist(suppu(1)n,suppu(2)n)+,lim infnRN(|un|2|u(1)n|2|u(2)n|2)dx0; (3.6)

    compactness: there exists a sequence {yn}RN with the following property: for any ε>0, there exists R>0 such that

    BR(yn)|un|2dxaε. (3.7)

    Claim 1. Vanishing does not occur.

    Otherwise, by Lemma I.1 of [38], we get un0 strongly in Lp(RN) for 2<p<2. Since 2<2(N+2+α)N+α<2, it follows from Lemma 2.2 that

    limnRN(IαF(un))F(un)dx=0.

    Then, by Lemma 3.1 and Eq (3.5) we have

    0>m(a)=limnIμ(un)=limn12|un|220.

    This contradiction proves Claim 1.

    Claim 2. Dichotomy does not occur.

    Otherwise, if dichotomy occurs, there exist b(0,a) and sequences {u(1)n}, {u(2)n} satisfying Eq (3.6). Furthermore, we may assume

    un=u(1)n+u(2)n+vn,u(1)nu(2)n=u(1)nvn=u(2)nvn=0almosteverywhereinRN.

    Then, we have

    |un|qq=|u(1)n|qq+|u(2)n|qq+|vn|qq, (3.8)

    and

    RN(IαF(un))F(un)dx=RN(IαF(u(1)n))F(u(1)n)dx+RN(IαF(u(2)n))F(u(2)n)dx+2RN(IαF(u(1)n))F(u(2)n)dx+2RN(IαF(un))F(vn)dx+RN(IαF(vn))F(vn)dx. (3.9)

    By (f1), (f2), and Lemma 2.1, we obtain

    |RN(IαF(un))F(vn)dx|C(|un|22+|un|2(N+2+α)N+α2(N+2+α)N+α)N+α2N(|vn|22+|vn|2(N+2+α)N+α2(N+2+α)N+α)N+α2N.

    By the definition of vn and Eq (3.6), we obtain that |vn|p0 for 2p<2. It follows from the above inequality and the boundedness of {un} that

    limnRN(IαF(un))F(vn)dx=0. (3.10)

    Using Eq (3.6) and Lemma 2.2, we get

    limnRN(IαF(vn))F(vn)dx=0. (3.11)

    Moreover, by the Young's inequality ([36] Theorem 4.2) and dist(suppu(1)n,suppu(2)n)+, we infer that

    limnRN(IαF(u(1)n))F(u(2)n)dx=0. (3.12)

    Thus, Eqs (3.6) and (3.8)–(3.12) imply that

    m(a)=limnIμ(un)lim supn(Iμ(u(1)n)+Iμ(u(2)n))m(b)+m(ab),

    which contradicts to Lemma 3.2, and proves Claim 2.

    Hence, the compactness holds, namely, there exists a subsequence {yn}RN such that ˜un=un(x+yn)˜u in L2(RN), ˜un˜u in H1(RN) and ˜uS(a). It follows from interpolation inequality and Sobolev inequality that

    limn|˜un˜u|p=0,for2<p<2.

    Then, Lemma 2.2 implies

    limnRN(IαF(˜un˜u))F(˜un˜u)dx=0.

    Hence,

    m(a)Iμ(˜u)lim infnIμ(˜un)=lim infnIμ(un)=m(a).

    Thus, we have Iμ(˜u)=m(a) and ˜un˜u as n. Moreover, un(x+yn)˜u in H1(RN).

    Proof. [Proof of Theorem 1.1] By Lemmas 3.1 and 3.3, there exists a global minimizer ˜u for Iμ on S(a) with m(a)=Iμ(˜u)<0. Furthermore, ˜u is a ground state solution of Eq (1.1) for some ˜λR. Then, by Lemma 2.4, we have

    P(˜u)=N22|˜u|22N2˜λ|˜u|22N+α2RN(IαF(˜u))F(˜u)dxNμq|˜u|qq=0.

    Then,

    λa=2m(a)2N|˜u|22αNRN(IαF(˜u))F(˜u)dx. (3.13)

    Since RN(IαF(˜u))F(˜u)dx>0, by Eq (3.13) we have ˜λ<0.

    By (f3), without loss of generality, we may assume that f0 on (0,+). Since f is odd, then F is even and thus for every uH1(RN), Iμ(|u|)=Iμ(u). From this one easily obtain that the function |˜u| is also a ground state solution of Eq (1.1). By regularity properties of [9], |˜u| is continuous, we can apply the strong maximum principle get |˜u|>0 on RN and thus ˜u has constant sign.

    Finally, we prove the symmetry of ˜u. Assume that HRN is a closed half-space and that σH denotes the reflection with respect to H. The polarization ˜uH(x):RNR of ˜u is defined for xRN by

    ˜uH(x):={max{˜u(x),˜u(σH(x))}ifxH,min{˜u(x),˜u(σH(x))}ifxH.

    By the properties of polarization ([9] Lemma 5.4), we observe that

    RN|˜uH|2dx=RN|˜u|2dx,

    and then

    RN|˜uH|sdx=RN|˜u|sdx,foranys[0,+).

    Moreover, since F is nondecreasing on (0,+), we have F(˜uH)=F(˜u)H. Therefore, by Lemma 5.5 of [9], we get

    Iμ(˜uH)Iμ(˜u) (3.14)

    with equality if and only if

    eitherF(˜u)H=F(˜u)orF(˜u)H=F(˜u(σH))inRN. (3.15)

    On the other hand, since ˜uS(a) and |˜u|22=|˜uH|22, we have Iμ(˜uH)m(a). It follows from Eq (3.14) that Iμ(˜uH)=Iμ(˜u), and thus Eq (3.15) holds. If F(˜u)H=F(˜u), for every xH,

    ˜u(x)˜u(σH(x))f(t)dt=F(˜u(x))F(˜u(σH(x)))=F(˜uH(x))F(˜u(σH(x)))0.

    Since F is nondecreasing on (0,+), we get ˜u(σH(x))˜u(x). In particular, f(˜uH)=f(˜u) on RN, hence ˜uH=˜u. If F(˜u(σH))=F(˜u)H, we similarly get ˜uH=˜u(σH). Since the hyperplane H is arbitrary, in either case we conclude that the function ˜u is radially symmetric with respect to some point x0RN, and is radially decreasing.

    In this section, we consider the case q=¯q and prove Theorems 1.2 and 1.3. Set ¯aN:=¯q2C¯qN,¯q>0.

    Lemma 4.1. Assume that (f1) and (f2) hold. For any μ>0, there exists a0(μ)=(¯aNμ1)N2, such that for any a(0,a0), we have

    <m(a)0.

    Proof. For every uS(a), by Eqs (2.2) and (3.1), we get

    Iμ(u)(12μC¯qN,¯q¯qa2N)|u|22C3C2εaN+2NC4ε2aα+2N|u|22.

    Since a<a0, we have 12μC¯qN,¯q¯qa2N>0. We choose ε>0 small such that C4ε2aα+2N=12(12μC¯qN,¯q¯qa2N), it follows that

    Iμ(u)12(12μC¯qN,¯q¯qa2N)|u|22CaN+2N.

    This implies m(a)>.

    In addition, for uS(a) we have

    m(a)Iμ(ut)=t2(12|u|22μ¯q|u|¯q¯q)12RN(IαF(ut))F(ut)dx. (4.1)

    Noticing that a<a0, by Eq (2.2), we have 12|u|22μ¯q|u|¯q¯q>0, and

    limt0t2(12|u|22μ¯q|u|¯q¯q)=0.

    Moreover, |ut|2(N+2+α)N+α2(N+2+α)N+α=t2NN+α|u|2(N+2+α)N+α2(N+2+α)N+α0 as t0. By Lemma 2.2, we have

    limt0RN(IαF(ut))F(ut)dx=0.

    Then, by Eq (4.1) we infer that

    m(a)limt0Iμ(ut)=0.

    Thus, m(a)0. The proof is completed.

    Next, we give some properties of m(a).

    Lemma 4.2. It holds that

    (i) Let a(0,a0). Then, we have for all b(0,a)

    m(a)m(b)+m(ab),

    and if m(b) or m(ab) is reached, then the inequality is strict.

    (ii) Taking μ>0 small enough, there exists b0>0 such that 0<b0<a0(μ). Then, for any a(b0,a0), we have m(a)<0.

    (iii) a(0,a0)m(a) is continuous.

    Proof. (i) Fix b(0,a), we first show that

    m(θb)θm(b),foranyθ(1,ab], (4.2)

    and that if m(b) is reached, the inequality is strict. By the definition of m(b), for any ε>0 sufficiently small, there exists a uS(b) such that

    Iμ(u)m(b)+ε. (4.3)

    Now set ¯u(x)=u(θ1Nx). Note that ¯u(x)S(θb). It follows from Eq (4.3) that

    m(θb)Iμ(¯u)<θIμ(u)θm(b)+θε. (4.4)

    Since ε>0 is arbitrary, we have that m(θb)θm(b). If m(b) is reached, we can let ε=0 in Eq (4.3), and thus Eq (4.4) implies m(θb)<θm(b).

    Then, by Eq (4.2) we have

    m(a)=abam(a)+bam(a)=abam(aab(ab))+bam(abb)m(ab)+m(b),

    with a strict inequality if m(b) is reached.

    (ii) By (f3), there exists vH1(RN) such that

    RN(IαF(v))F(v)dx>0.

    For any b>0, set vb(x)=v(b1N|v|2N2x). Obviously vbS(b). Then, we have

    Iμ(vb)=12RN|vb|2dx12RN(IαF(vb))F(vb)dxμ¯qRN|vb|¯qdx=bN2N2|v|2(N2)N2|v|22bN+αN2|v|2(N+α)N2RN(IαF(v))F(v)dxμb¯q|v|22RN|v|¯qdx:=g(b).

    Since g(b) as b+ and by choosing b0>0 large such that Iμ(vb0)<0, it follows that

    m(b0)Iμ(vb0)<0.

    Now, taking μ>0 small enough such that b0<a0(μ). For any a(b0,a0), by Lemma 4.1 and (i), we obtain

    m(a)m(ab0)+m(b0)m(b0)<0.

    (iii) Let a(0,a0) be arbitrary and {an}(0,a0) be such that ana as n. By the definition of m(an), for every n there exists unS(an) such that

    Iμ(un)m(an)+1n. (4.5)

    Since m(an)0, by the proof of Lemma 4.1, the sequence {un} is bounded in H1(RN). Set vn=aanun. It is clear that vnS(a). Then, we can write

    m(a)Iμ(vn)=Iμ(un)+(Iμ(vn)Iμ(un)), (4.6)

    where

    Iμ(vn)Iμ(un)=12(aan1)|un|22μ¯q[(aan)¯q21]|un|¯q¯q12RN[(IαF(vn))F(vn)(IαF(un))F(un)]dx. (4.7)

    By the boundedness of {un}, we have {|un|¯q¯q} is bounded. Thus, we have

    limn(aan1)|un|22=0=limn[(aan)¯q21]|un|¯q¯q. (4.8)

    Moreover,

    RN[(IαF(vn))F(vn)(IαF(un))F(un)]dx=RN(IαF(vn))(F(vn)F(un))dx+RN(Iα(F(vn)F(un)))F(un)dx. (4.9)

    Since ana as n, then {aan} is bounded. It follows from (f1) and (f2) that

    |F(vn)F(un)|10|f(un+t(vnun))||vnun|dtC(aan1)(|un|N+αN+|un|N+α+2N).

    Thus, using Lemma 2.1 and the Sobolev imbedding inequality, we get

    |RN(Iα(F(vn)F(un)))F(un)dx|Cα|F(un)|2NN+α|F(vn)F(un)|2NN+αC(aan1)(|un|22+|un|2(N+α+2)N+α2(N+α+2)N+α)N+αN.

    Thus,

    limnRN(Iα(F(vn)F(un)))F(un)dx=0.

    Similarly,

    limnRN(IαF(vn))(F(vn)F(un))dx=0.

    Then, in view of Eq (4.9), we have

    limnRN[(IαF(vn))F(vn)(IαF(un))F(un)]dx=0. (4.10)

    By Eqs (4.7), (4.8), and (4.10), we have

    limn(Iμ(vn)Iμ(un))=0. (4.11)

    It follows from Eqs (4.5) and (4.6) that

    m(a)lim infnm(an).

    On the other hand, for any ε>0, there exists uS(a) such that

    Iμ(u)m(a)+ε. (4.12)

    Set ˜un(x)=anau(x). Then, ˜unS(an). Similar to Eq (4.11), we have

    Iμ(˜un)=Iμ(u)+on(1).

    It follows from Eq (4.12) that

    m(an)Iμ(˜un)=Iμ(u)+(Iμ(˜un)Iμ(u))m(a)+ε+on(1).

    Since ε>0 is arbitrary, we get

    lim supnm(an)m(a).

    Thus, we infer that m(an)m(a) as n.

    Now, for any fixed μ>0 small enough, we define

    a=inf{a:0<a<a0(μ),m(a)<0}.

    By Lemma 4.1, a[0,a0(μ)) is well defined and satisfying

    m(a)=0if0<aa,m(a)<0ifa<a<a0. (4.13)

    Lemma 4.3. Assume that (f1)–(f3) hold. For any a(0,a0), if (f4) holds, then a=0; if (f5) holds, then a>0.

    Proof. (i) Let a(0,a0), we choose uS(a)C0(RN) and set

    L=(2|u|22RN(Iα|u|N+2+αN)|u|N+2+αNdx)12>0.

    By the assumption (f4), there exists δ>0 such that

    F(s)L|s|N+2+αNforall|s|<δ.

    Since |ut|δ for sufficiently small t>0, we have

    Iμ(ut)12|ut|2212RN(IαF(ut))F(ut)dx12t2|u|22L22t2RN(Iα|u|N+2+αN)|u|N+2+αNdx=12t2|u|22<0.

    Thus, m(a)Iμ(ut)<0 for any a(0,a0). By the definition of a, we get a=0.

    (ii) By (f2) and (f5), there exists C>0 such that

    |F(s)|C|s|N+2+αNforallsR. (4.14)

    For uS(a), by Eqs (2.1) and (4.14), we get

    RN(IαF(u))F(u)dxCa2+αN|u|22.

    It follows from Eq (2.2) that

    Iμ(u)12|u|22C2a2+αN|u|22μ¯qC¯qN,¯qa2N|u|22.

    Taking a>0 small enough such that C2a2+αN+μ¯qC¯qN,¯qa2N12, we get Iμ(u)0. This implies m(a)0 for a small a>0. From Lemma 4.1 that m(a)=0 for a small a>0. Hence, we have a>0.

    Lemma 4.4. Assume that (f1)(f3) hold. Let μ>0 and a(0,a0). Let {un}S(a) be a minimizing sequence for m(a). Then, one of the following holds:

    (i).

    lim supnsupzRNB1(z)|un|2dx=0.

    (ii). Taking a subsequence if necessary, there exist uS(a) and a family {yn}RN such that un(+yn)u in H1(RN) as n. Specifically, u is a global minimizer.

    Proof. Suppose that {un}S(a) is a minimizing sequence which does not satisfy (i). Then, we have

    0<limnsupzRNB1(z)|un|2dxa.

    Taking a subsequence if necessary, we may assume there exists a family {yn}RN such that

    0<limnB1(yn)|un|2dxa.

    Let us consider un(+yn). Since {un}S(a) is a minimizing sequence of m(a), by the proof of Lemma 4.1, {un} is bounded in H1(RN), Hence, {un(+yn)} is bounded in H1(RN). Then, there exists uH1(RN) such that un(+yn)u in H1(RN), un(+yn)u in Lploc(RN) for 2p<2, and un(+yn)u almost everywhere in RN. It follows that

    B1(0)|un(x+yn)|2dx>0.

    Then, by un(+yn)u in L2loc(RN), we obtain that u0 and |u|22>0. Set vn(x)=un(x+yn)u. Then, vn0 in H1(RN), and we have

    |un(+yn)|22=|un|22=|vn+u|22=|vn|22+|u|22+on(1), (4.15)
    |un(+yn)|22=|un|22=|(vn+u)|22=|vn|22+|u|22+on(1), (4.16)
    |un(+yn)|¯q¯q=|un|¯q¯q=|vn+u|¯q¯q=|vn|¯q¯q+|u|¯q¯q+on(1). (4.17)

    Moreover, by Lemma 2.3, we have

    RN(IαF(un))F(un)dx=RN(IαF(vn))F(vn)dx+RN(IαF(u))F(u)dx+on(1). (4.18)

    Since Iμ(un)=Iμ(un(+yn))=Iμ(vn+u), by Eqs (4.16)–(4.18), we obtain

    Iμ(un)=Iμ(vn)+Iμ(u)+on(1). (4.19)

    Claim. limn|vn|22=0.

    Denote b=|u|22>0. By Eq (4.15), we see that ba, and if we prove that b=a, then the claim holds. We suppose by contradiction that b<a. Since vnS(|vn|22), then

    Iμ(vn)m(|vn|22).

    It follows from Iμ(un)m(a) and Eq (4.19) that

    m(a)=Iμ(vn)+Iμ(u)+on(1)m(|vn|22)+Iμ(u)+on(1).

    By Lemma 4.2 (iii) and Eq (4.15), we obtain

    m(a)m(ab)+Iμ(u). (4.20)

    Note that b=|u|22, we have uS(b) and Iμ(u)m(b). If Iμ(u)>m(b), then Lemma 4.2 (i) and Eq (4.20) imply

    m(a)>m(ab)+m(b)m(a),

    which is impossible. Hence, Iμ(u)=m(b), that is u is a minimizer of m(b) on S(b). It follows from Lemma 4.2 (i) and Eq (4.20) that

    m(a)m(ab)+Iμ(u)=m(ab)+m(b)>m(ab+b)=m(a),

    which is a contradiction. Thus, the claim holds and Eq (4.15) implies |u|22=a.

    Finally, we show that limn|vn|22=0. Since uS(a), we have Iμ(u)m(a). It follows from Eq (4.19) that

    m(a)=limnIμ(vn)+Iμ(u)limnIμ(vn)+m(a),

    this implies

    limnIμ(vn)0. (4.21)

    On the other hand, by Lemma 2.2, Eq (2.2), and |vn|220, we infer that

    limnRN(IαF(vn))F(vn)dx=0,limnRN|vn|¯qdx=0.

    In view of Eq (4.21), we get

    limn12|vn|22=limnIμ(vn)0.

    Then, we have limn|vn|22=0. It follows from |vn|220 that vn0 in H1(RN). Hence, un(+yn)u in H1(RN) as n.

    Proof. [Proof of Theorem 1.2] (i) For the case 0<a<a, we assume by contradiction that there exists a global minimizer with respect to m(a). By Eq (4.13), we know m(a)=0. By Lemma 4.2 (i) with the strict inequality, we infer that

    0=m(a)<m(aa)+m(a)=m(a)=0,

    which is a contradiction.

    (ii) For the case a<a<a0. By Eq (4.13), we have m(a)<0. Let {un}S(a) satisfying limnIμ(un)=m(a). It is easy to see that {un} is bounded in H1(RN). Then, one of the two alternatives in Lemma 4.4 occurs. Let assume that (i) of Lemma 4.4 take place. The Lemma 1.1 of [38] implies un0 in Lt(RN) for 2<t<2. Moreover, since 2<2(N+2+α)N+2<2, by Lemma 2.2 we have

    limnRN(IαF(un))F(un)dx=0andlimnRN|un|¯qdx=0.

    From this, we infer that

    m(a)=limnIμ(un)=limn12|un|220,

    which contradicts to m(a)<0. Thus, Lemma 4.4 (ii) holds, namely, there exist uS(a) and a family {yn}RN such that un(+yn)u in H1(RN) as n, and u is a global minimizer to m(a). Moreover, u is a ground state solution to Eq (1.1) for some λR. By Lemma 2.4, Eq (3.13), and m(a)<0, we have λ<0. Similar as the proof of Theorem 1.1, u has constant sign and is radially symmetric about a point in RN.

    Proof. [Proof of Theorem 1.4] By (f6) and (iii) of Theorem 1.2, we have a>0. Let an=a+ln for any nN, where 0<l<a0a. By Eq (4.13) and Theorem 1.2 (ii), for every n, there exists a global minimizer unS(an) such that

    Iμ(un)=m(an)<0. (4.22)

    Then, {un} is bounded in H1(RN). Since ana, we have limn|un|22=a. Set vn=a|un|2un. Clearly, vnS(a) and a|un|21 as n. By the similar proof of Lemma 4.2 (iii), we obtain

    limnIμ(vn)=limnIμ(un)=limnm(an)=m(a)=0.

    Thus, {vn} is a minimizing sequence with respect to m(a).

    Next, we prove that {vn} is non-vanishing, that is

    δ:=limnsupzRNB1(z)|vn|2dx>0.

    Otherwise, δ=0. By the definition of vn and a|un|21, we get

    limnsupzRNB(z,1)|un|2dx=0. (4.23)

    By (ii) of Theorem 1.2, un satisfies Eq (1.1) and we may assume that un is radially symmetric with respect to the origin and decreasing for any nN. Using the elliptic regularity theory, we see that {un} is bounded in C1(B(0,1)). Thus, by Eq (4.23) we get

    un(0)=|un|0asn.

    It follows from (f6) that for any ε>0 and sufficiently large n,

    RN(IαF(un))F(un)dxCε2a2+αNn|un|22Cε2(a+l)2+αN|un|22.

    Choosing ε>0 small enough such that Cε2(a+l)2+αN12μ¯qC¯qN,¯q(a+l)2N, then by Eq (2.2) we have

    Iμ(un)12(12μ¯qC¯qN,¯q(a+l)2N)|un|220,

    which contradicts Eq (4.22), and hence δ>0.

    Then Lemma 4.4 (ii) holds, that is, up to a subsequence, there exist vS(a) and a family {yn}RN such that vn(+yn)v in H1(RN) as n. Then, v is a global minimizer to m(a). Moreover, v is a ground state solution to (1.1) for some λR. By Lemma 2.4 and Eq (3.13), we have

    λa=2m(a)2N|v|22αNRN(IαF(v))F(v)dx=2N|v|22αNRN(IαF(v))F(v)dx<0,

    which implies λ<0. By the same arguments of Theorem 1.1, we obtain the symmetric of v.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors would like to express sincere thanks to the anonymous referees for their carefully reading this paper and valuable useful comments. This work was supported by Qinglan Project of Jiangsu Province (2022).

    The authors declare there is no conflicts of interest.

    [1] Moghadam AD, Schultz BF, Ferguson JB, et al. (2014) Functional metal matrix composites: self-lubricating, self-healing, and nanocomposites-an outlook. JOM 66: 872–881.
    [2] Sato A, Mehrabian R (1976) Aluminum matrix composites: fabrication and properties. Metall Trans B 7: 443–451.
    [3] Amirkhanlou S, Niroumand B (2010) Synthesis and characterization of 356-SiCp composites by stir casting and compocasting methods. Trans Nonferrous Met Soc China 20: s788–s793. doi: 10.1016/S1003-6326(10)60582-1
    [4] Das S, Das K (2007) Abrasive wear of zircon sand and alumina reinforced Al–4.5 wt% Cu alloy matrix composites–A comparative study. Compos Sci Technol 67: 746–751.
    [5] Rohatgi PK, Schultz BF, Daoud A, et al. (2010) Tribological performance of A206 aluminum alloy containing silica sand particles. Tribol Int 43: 455–466.
    [6] Dorri Moghadam A, Ferguson JB, Schultz BF, Lopez HF, Rohatgi PK (2016) Direct synthesis of nano structured in-situ hybrid aluminum matrix nanocomposite.Ind Eng Chem Res 55: 6345–6353.
    [7] Zuhailawati H, Samayamutthirian P, Haizu CM (2007) Fabrication of low cost aluminium matrix composite reinforced with silica sand. J Phys Sci 18: 47–55.
    [8] Yoshikawa N, Kikuchi A, Taniguchi S (2002) Anomalous temperature dependence of the growth rate of the reaction layer between silica and molten aluminum. J Am Ceram Soc 85: 1827–1834.
    [9] Hemanth J (2009) Quartz (SiO 2p) reinforced chilled metal matrix composite (CMMC) for automotive applications. Mater Des 30: 323–329. doi: 10.1016/j.matdes.2008.04.064
    [10] Sulaiman S, Sayuti M, Samin R (2008) Mechanical properties of the as-cast quartz particulate reinforced LM6 alloy matrix composites. J Mater Process Technol 201: 731–735. doi: 10.1016/j.jmatprotec.2007.11.221
    [11] Rohatgi PK, Pai BC, Panda SC (1979) Preparation of Cast Aluminum-Silica Particulate Composites. J Mater Sci 14: 2277–2283.
    [12] Rohatgi PK, Asthana R, Das S (1986) Solidification, structures, and properties of cast metal-ceramic particle composites. Int Mater Rev 31: 115–139. doi: 10.1179/imr.1986.31.1.115
    [13] Gupta AK, Dan TK, Rohatgi PK (1986) Aluminum Alloy-silica Sand Composites: Preparation and Properties. J Mater Sci 21: 3413–3419. doi: 10.1007/BF02402980
    [14] Moghadam AD, Omrani E, Menezes P L, Rohatgi PK (2016). Effect of in-situ processing parameters on the mechanical and tribological properties of self-lubricating hybrid aluminum nanocomposites.Tribology Letters62: 1-10.
    [15] Pai BC, Ramani G, Pillai RM, et al. (1995) Role of Magnesium in Cast Aluminum Alloy Matrix Composites. J Mater Sci 30: 1903–1911.
    [16] McLeod AD, Gabryel CM (1992) Kinetics of the Growth of Spinel. MgAl2O4, on Alumina Particulate in Aluminum Alloys Containing Magnesium. Metall Trans A 23A: 1279–1283.
    [17] Mogilevsky R, Bryan SR, Wolbach WS, et al. (1995) Reactions at the Matrix/Reinforcement Interface in Aluminum Alloy Matrix Composites. Mater Sci Eng A 191: 209–222. doi: 10.1016/0921-5093(94)09635-A
    [18] Hanabe MR, Aswath PB (1996) Al2O3/Al particle-reinforced aluminum matrix composite by displacement reaction. J Mater Res 11: 1562–1569. doi: 10.1557/JMR.1996.0195
  • Reader Comments
  • © 2016 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(5762) PDF downloads(1266) Cited by(7)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog