Content area

Abstract

In this paper, we study a class of nonlocal Schrödinger–Poisson–Slater equations: Δu+u+λIα|u|q|u|q2u=|u|p2u, where q,p>1, λ>0, and Iα is the Riesz potential. We obtain the existence, stability, and symmetry-breaking of solutions for both radial and nonradial cases. In the radial case, we use variational methods to establish the coercivity and weak lower semicontinuity of the energy functional, ensuring the existence of a positive solution when p is below a critical threshold p¯=4q+2α2+α. In addition, we prove that the energy functional attains a minimum, guaranteeing the existence of a ground-state solution under specific conditions on the parameters. We also apply the Pohozaev identity to identify parameter regimes where only the trivial solution is possible. In the nonradial case, we use the Nehari manifold method to prove the existence of ground-state solutions, analyze symmetry-breaking by studying the behavior of the energy functional and identifying the parameter regimes in the nonradial case, and apply concentration-compactness methods to prove the global well-posedness of the Cauchy problem and demonstrate the orbital stability of the ground state. Our results demonstrate the stability of solutions in both radial and nonradial cases, identifying critical parameter regimes for stability and instability. This work enhances our understanding of the role of nonlocal interactions in symmetry-breaking and stability, while extending existing theories to multiparameter and higher-dimensional settings in the Schrödinger–Poisson–Slater model.

Full text

Turn on search term navigation

1. Introduction

We study the nonlocal Schrödinger–Poisson–Slater (SPS)-type equation

(1)iψtΔψψ+λIα|ψ|q|ψ|q2ψ=|ψ|p2ψ,

where ψ:RN×(0,)C is the complex-valued wave function, p,q>1 are the exponents of the local and nonlocal nonlinearities, respectively, λR is the coupling constant that measures the strength of the long-range interaction, and Iα is the Riesz potential of order α(0,N), defined for xRN\{0} as

Iα(x)=Aα|x|Nα,Aα=Γ(Nα2)Γ(α2)πN/22α.

The normalization constant Aα is chosen so that the kernel Iα satisfies the semigroup property:

Iα+β=IαIβforallα,β(0,N)suchthatα+β<N.

For a detailed discussion, see [1]. The Riesz potential Iα in our nonlocal term is intimately connected to fractional Laplacian operators through their shared scaling properties and Fourier multipliers. In fact, just as the fractional Laplacian is a nonlocal operator defined via the Fourier transform with symbol |ξ|α, the convolution Iα|u|q embodies a similar nonlocal behavior. This connection motivates the use of fractional or Coulomb–Sobolev spaces to capture the appropriate function space setting, as detailed in the works of Moroz and van Schaftingen [2] and Di Nezza et al. [3]. The nonlocal convolution term in our equation represents the Coulombic repulsion between electrons. The local term |u|q2u was introduced by Slater [4] with q=83 as a local approximation of the exchange potential in the Hartree–Fock model [5,6].

In particular, we focus on solitary wave solutions, i.e., standing waves of the form

ψ(x,t)=eiωtu(x),

where we work in atomic units and shift the reference energy so that the coefficient of the linear term is normalised to 1; this convention fixes ω=2, which reduces the time-dependent equation to the stationary equation with ω=2 (see also [2])

(2)Δu+u+λ(Iα|u|q)|u|q2u=|u|p2u,uRN,

where NN,p>1,q>1, λ>0. Equation (2) thus represents a fractional Schrödinger–Poisson–Slater model in which the local power nonlinearity competes with the nonlocal convolution term, a setting widely used to analyze solitary-wave dynamics.

The natural working space for understanding Equation (2) is the space

Eα,q=uD1,2RN:RN×RN|u(x)|q|u(y)|q|xy|Nαdxdy<+.

The associated functional for the problem (2) is

E(u)=12RN|u|2+|u|2dx+λ2qRN×RN|u(x)|q|u(y)|q|xy|Nαdxdy1pRN|u|pdx

and Equation (2) is the Euler–Lagrange equation of the functional E(u) on the space Eα,qL2.

To guarantee the well-definedness of the functional, we define the space

Eα,qL2(RN)

with the norm

uEα,qL2(RN)=uEα,q+uL2,

which makes (Eα,qL2(RN),·X) a Banach space. The norm ensures that the functional E(u) is finite for all uEα,qL2(RN). Equivalently, Eα,qL2(RN) has the product topology induced by the norm ·Eα,q+·L2, making both projections continuous. A sequence unu in Eα,qL2(RN) if and only if unu in both Eα,q and L2.

For more regularity in the solutions, we consider the space Eα,qH1RN. This space is the intersection of the Coulomb–Sobolev space Eα,qRN and the Sobolev space H1RN, which is defined as

Eα,qH1RN=uD1,2RN:RN×RN|u(x)|q|u(y)|q|xy|Nαdxdy<+,uH1RN.

This space ensures that the solutions have the required regularity, with square-integrable gradients (i.e., uH1(RN)), while also satisfying the nonlocal integrability condition. It is particularly useful for analyzing the stability and behavior of solutions, especially in the presence of nonradial symmetry.

The stationary Equation (2) can be derived as the Euler–Lagrange equation of the Thomas–Fermi–Dirac–von Weizsöcker (TFDW) energy, which itself originates from the Hartree–Fock self-consistent field method by replacing the nonlocal exchange integral with a local power-law term and adding a gradient correction to the Thomas–Fermi kinetic energy and a Dirac exchange correction, yielding the Coulomb–Dirichlet functional

RN|u|2dx+RN×RN|u(x)|2|u(y)|2|xy|Nαdxdy,

in the absence of external potentials [7,8]. Density functional theory (DFT) expresses the total energy of a quantum system purely in terms of its electron density. In the local-density approximation, this framework recovers the energy functional whose Euler–Lagrange equation becomes the Schrödinger Poisson Slater (SPS) model [9,10,11], thereby unifying long-range Coulomb interactions with a computationally tractable local nonlinear exchange term. Rigorous variational analysis of (2) hinges on two key points: the Hardy–Littlewood–Sobolev (HLS) inequality [12,13], which bounds the nonlocal convolution term, and concentration-compactness methods, which address the lack of compactness in the Coulomb–Sobolev space to establish critical embeddings. Moreover, when q=2 and α=N2, (2) reduces to the classical Choquard equation whose ground-state theory was pioneered by [2,14,15]. Also, when N=3, q=2 and p=6, Jeanjean and Le [16] studied Equation (2) does not admit positive solutions.

With the change of variable v(x)=ε2p2u(εx), λ=ε4q4p2α2, we convert the problem (2) to

(3)Δu+ε2u+Iα|u|q|u|q2u=|u|p2u.

The parameter ε controls the scaling of the solution and spatial variables, enabling asymptotic analysis by homogenizing different scales in the system. Motivated by this equation, we study the limit problem first

(4)Δu+Iα|u|q|u|q2u=|u|p2u.

Throughout this paper, we assume that λ>0; hence, we omit its explicit mention in subsequent sections. According to [17], we can also say that it is a “zero-mass” problem, since the zero linearization operator only involves the Laplacian. The associated energy functional is J:ErdR,

J(v)=12RN|v|2dx+12qRN×RN|v(x)|q|v(y)|q|xy|Nαdxdy1pRN|v|pdx.

It is well-established that J(v) is well-defined, and C1 and its critical point are associated with the solution of (4).

Our goal in this paper is to establish the existence of ground-state solutions and to prove their orbital stability for the time-dependent problem (1), and to investigate the qualitative properties of solutions to the associated stationary problem (2).

For a classical case in (2), see the Schrödinger–Poisson–Slater problem:

Δu+u+λu21|x|u=|u|p2u

where uH1R3, λ>0 and p(2,3). Ruiz [18] demonstrates a lower bound for the Coulomb energy and proposes an inequality that is versatile across different frameworks. Importantly, the derived inequality is nearly optimal, underlining its close proximity to optimality within the specified context.

Ruiz’s approach for controlling the Riesz potential fails when q2 because it relies on the specific structure of the Coulomb-type convolution, which is valid only for q=2. In this case, the nonlocal term corresponds to the Coulomb kernel, enabling the use of symmetry-based compactness arguments. However, for q2, the convolution becomes more complex and loses this structure, complicating the analysis due to the altered interaction between nonlocal and nonlinear terms. Consequently, it is crucial to carefully examine how p and q interact in order to properly balance the effects of nonlocality and nonlinearity. This requires developing new mathematical techniques that go beyond the traditional Coulomb framework.

In [19], Mercuri et al. studied the nonlocal Schrödinger–Poisson–Slater equation

Δu+Iα|u|p|u|p2u=|u|q2uinRN,

with NN, p>1, q>1. They introduced the Coulomb–Sobolev space and established optimal interpolation inequalities that yield existence results in specific parameter regimes. Their embedding results have been particularly influential in our study.

We allow arbitrary dimension N3, Riesz potential order α(0,N), and convolution exponent q1, thereby generalizing the problem beyond the constraints considered by Mercuri et al. [19]. By incorporating the linear mass term, we unify the analysis of zero-frequency and nonzero-frequency cases, capturing the frequency’s impact on ground state existence and stability, a factor not addressed in Ruiz’s bounded domain analysis [18]. Also, in contrast to Bellazzini et al. [20], who mainly focused on the existence of standing waves under an L2 mass constraint, we eliminate this constraint using the linear term, making the energy functional coercive. We prove the orbital stability of radial ground states when p lies below the critical threshold p¯=4q+2α2+α, and establish nonexistence when p exceeds the Sobolev critical exponent 2*=2NN2 using the Pohozaev identity. This precise delineation of the supercritical regime, which previous studies either partially addressed or overlooked [19,20].

Moreover, the framework developed by Moroz and Van Schaftingen [2] for fractional Choquard equations, along with insights from Servadei and Valdinoci [21], underscores the importance of nonlocal variational methods. Recent contributions by Tang et al. [22] and Liu et al. [23] further demonstrate existence, nonexistence, and symmetry-breaking phenomena in related models, albeit within restricted parameter ranges and dimensions. These developments collectively motivate our broader investigation into fractional nonlocal Schrödinger–Poisson–Slater equations. Notably, Zhang and Hou [24] and Wang et al. [25] have provided crucial insights into the existence and radial symmetry of ground states, while Zhang et al. [26] and Zhang et al. [27] advanced the analysis of fractional p-Laplacian systems. Furthermore, recent work by Zhang and Nie [28] and by Wang et al. [29] has introduced novel tools for fractional operators that are instrumental in addressing symmetry-breaking in the nonradial setting.

It is noteworthy that our results have broad implications; indeed, several existing works (e.g., [18,23,30,31,32]) appear as particular cases when the parameters ε,N,p,q, and α are suitably specialized. By refining the Coulomb–Sobolev embedding of Mercuri et al. [19], we prove convergence and existence across a substantially wider parameter domain-recovering Ruiz’s coercivity for q=2, p(18/7,3) [18]-and further extending their framework to supercritical p and nonradial minimisers via a nonlocal Brezis–Lieb lemma (see Theorem 1 and Section 4).

Throughout this paper, we fix

N3,0<α<N,p>1,q>1,λ>0,

and set

2*=2NN2,p¯=4q+2α2+α,prd=q+Nqαq+2N2α3N+α4.

And Erd represents the subspace of radial functions in the space Eα,q.

We have the following result:

Theorem 1. 

Let N3 and 0<α<N. assume either

p ( p r d , p ¯ ) with q < N + α N 2 or p ( 2 * , p ¯ ) with q > N + α N 2 .

Then, the functional

J : E r d R

is coercive and weakly lower semicontinuous. In particular, if p<p¯, J attains a negative minimum, so that Equation (4) admits a positive solution in Erd.

Our approach extends Theorem 1.3 in [18], which proved coercivity and minimizers for the radial Schrödinger–Poisson–Slater functional J with p(18/7,3) in R3, by generalizing to multiparameter dependencies, regime bifurcations involving both cases q>N+αN2 and q<N+αN2, and higher-dimensional settings. This requires overcoming new challenges in scaling, critical exponents, and nonlocal decay.

Remark 1. 

In the special case q=2,α=2, and N=3, Theorem 1.3 of [18] recovers a similar result. Our work broadens this framework, allowing exponents p and q to exceed the critical threshold 2*. Notably, our proof leverages the compact immersion from [20], enabling this extension to supercritical regimes.

Theorem 2. 

Let N3,0<α<N. Assume that

p p r d , 2 * , q < N + α N 2 .

Then, the functional E(u) associated with Equation (2), is coercive, weakly lower semicontinuous, attains a minimum, and satisfies the Palais–Smale condition.

After establishing the existence of ground states, we now derive the Pohozaev identity to pinpoint the parameter regimes that rule out nontrivial solutions.

Theorem 3 

(Nonexistence of Ground States). Let N3 and 0<α<N. Assume that

p>2NN2andq<N+α2Np.

Then, any solution uEα,qL2 of Equation (2) must be trivial, i.e., u0.

In nonlocal Schrödinger–Poisson–Slater equations, symmetry-breaking can occur when the nonlocal interactions, represented by the Riesz potential, dominate over local terms. Mercuri et al. [19] analyze this competition and show how the nonlocal term affects the symmetry of solutions, particularly in regimes where nonlocality becomes dominant. Using Coulomb–Sobolev spaces and Brezis–Lieb-type inequalities, they explore the existence and properties of ground states, highlighting how the symmetry of solutions can break depending on the relative strengths of the local and nonlocal terms. Liu et al. [23] further explored this, identifying both radial and nonradial minimizers, while Ruiz [18] highlighted the conditions that lead to symmetry-breaking in the SPS system.

In the following, we demonstrate that the interplay between the nonlocal and nonlinear terms governs the existence, with the choice and interplay of parameter regimes playing a decisive role.

Theorem 4. 

Let N3 and α(0,N). Assume that

(5) 2 ( 2 q + 2 N α ) 2 N + 2 α < p < 2 N N 2 , q < N + α N 2 .

Then, there exists a ground-state solution u to Equation (2) that is orbitally stable.

We will write some common symbols as follows:

Cc(RN) is the space of compactly supported C-functions.

H1(RN) is endowed with the norm

uH12=RN|u|2+|u|2dx.

D1,2(RN) is endowed with the norm

uD1,22=RN|u|2dx.

We write f(t)g(t) as t if limtf(t)/g(t)=1.

To provide a clear overview of the main results discussed in this paper, we present a summary of the parameter ranges and corresponding conclusions of the four primary theorems in Table 1.

The paper is organized as follows: Section 2 presents the variational framework, functional spaces (including Erd), and preliminary tools. Section 3 proves the coercivity, existence and nonexistence of minimizers for J and E (Theorems 1–3). Section 4 establishes the existence and orbital stability of ground states (Theorem 4).

2. Preliminaries

In this section, we establish some notation that will be used throughout the article. We also study some basic properties of the space Eα,q.

Definition 1. 

The space D1,2RN is defined as

D 1 , 2 R N = u L 2 * R N : | u | L 2 R N

with the norm

u D 1 , 2 = R N | u ( x ) | 2 d x 1 / 2 ,

and define the space Eα,qRN

E α , q = u D 1 , 2 R N : R N | u | 2 d x + R N × R N | u ( x ) | q | u ( y ) | q | x y | N α d x d y <

with the norm of Eα,qRN

u E α , q R N = R N | u ( x ) | 2 d x + R N × R N | u ( x ) | q | u ( y ) | q | x y | N α d x d y 1 / q 1 / 2

As shown in [19] (see also [20,33]), the space Es,α,q(RN) is a normed space, and it is also complete. That is to say, the space Eα,q(RN) is a Banach space (see Proposition 1 below). For convenience, we state it below for our later use.

Proposition 1. 

For every NN,0<α<N and q[1,), the normed space E(RN) is a Banach space, i.e., ·E is a norm of E, and E,·E is a uniformly convex Banach space. Moreover, it is a complete normed space.

We shall use the following result, which is also from [19,20]. The Coulomb–Sobolev space E can be naturally approximated in the ·E norm by the set of testing functions Cc(RN), which is proved in [19]. For the sake of readers, we recall its statement here.

Proposition 2 

(Density of test functions). Let NN, α(0,N) and p1. Then, Cc(RN) is dense in E.

We adapt the following compact embedding result from [19,20], adjusted to our parameter framework:

Lemma 1. 

Let N2, q[1,), and 1<α<N. Assume p satisfies:

p 2 * , p r d i f 1 q < N 2 N + α , p p r d , 2 * i f 1 q > N 2 N + α ,

where 2*=2NN2 is the critical Sobolev exponent and prd=q+Nqαq+2N2α3N+α4. Then, the embedding

E r d ( R N ) L p ( R N )

is compact.

We employ a scaling-invariant inequality from [20], adjusted to our parameter regime. Specifically, for our choice of α,q,N, and p, the following holds:

Lemma 2 

(Scaling–Invariant Sobolev–Coulomb inequality). Let NN, 0<α<N, and p,qR satisfy q(N2)N+α. There exists a constant C=C(N,α,q,p)>0 such that the scaling-invariant inequality

ϕLp(RN)CϕD1,2(RN)p(N+α)2Nqp(N+α2)RN×RN|ϕ(x)|q|ϕ(y)|q|xy|Nαdxdy2Np(N2)2p(N+αq(N2))

holds for every ϕEα,q(RN) if and only if

pp¯,2*if1q>N2N+α,p2*,p¯if1q<N2N+α.

Proposition 3 

(Sobolev embedding). Let N3. Sobolev space D1,2RN continuously embeds into L2*(RN). Specifically, there exists a constant C=C(N)>0 such that:

uL2*RNCuL2RNuD1,2RN.

Lemma 3. 

For λ>0, the nonlocal term Iα|u|q|u|qdx modifies the mountain-pass geometry.

Proof. 

Recall the energy functional associated with (2):

E(u)=12|u|2+|u|2dx+λ2qIα|u|q|u|qdx|u|pdx.

Now, fix a nonzero u. Define

γ(t)=E(tu)=t22(uL22+uL22)+t2q2qλIα|u|q|u|qdxtp|u|pdx,t>0.

As t0, we have

γ(t)=t22(uL22+uL22)+ot2

so that γ(t)>0 for sufficiently small t. On the other hand, as t, the local term tp|u|pdx dominates provided p>2q. In fact,

γ(t)tp|u|pdx<0

for t sufficiently large. Next, differentiating γ(t) with respect to t, we obtain

γ(t)=t(uL22+uL22)+t2q1λIα|u|q|u|qdxptp1|u|pdx.

Setting γ(t)=0 for t>0 and dividing by t, we obtain

(uL22+uL22)+t2q2λIα|u|q|u|qdx=ptp2|u|pdx.

Next, differentiating twice with respect to t yields

γ(t)=uL22+uL22+(2q1)t2q2λIα|u|q|u|qdxp(p1)tp2|u|pdx

When q<N+αN2 and pprd,p¯, and when q>N+αN2 and p2*,p¯, we have that for t0, the quadratic term dominates so that γ(t)>0; whereas for t, the negative term p(p1)tp2|u|pdx eventually dominates, and hence, γ(t)<0. This change in sign implies that γ(t) has a unique positive maximum at some t=t0.

Hence, by continuity of γ(t), there is at least one positive solution t0>0 to γt0=0. We claim the uniqueness of positive solution t0. Suppose, for contradiction, there exist two positive roots 0<t1<t2 of γ(t)=0. By continuity, γ(t) must change sign at t1, becoming either strictly positive or negative for t>t1. However, we already know that γ(t) as t. Hence, γ cannot become positive again after t1, ruling out a second zero at t2.

Therefore, the energy functional E(u) possesses the mountain-pass geometry. □

To establish various key estimates in the proof, we rely on the well-known Hardy–Littlewood–Sobolev inequality, which plays a crucial role in controlling nonlocal terms and ensuring the coercivity of the energy functional [13].

Lemma 4 

([13]). Let p,r>1 and 0<λ<N with 1/p+λ/N+1/r=2. Let fLpRN and hLrRN. Then, there exists a sharp constant C(N,λ,p), independent of f and h, such that

(6)RNRNf(x)|xy|λh(y)dxdyC(N,λ,p)fphr.

The sharp constant satisfies

C(N,λ,p)N(Nλ)SN1/Nλ/N1prλ/N11/pλ/N+λ/N11/rλ/N.

3. The Existence of the Solution in the Radial Case

In this section, we establish three results, Theorems 1–3, in the radial setting. The precise parameter ranges play a critical role in our analysis.

Proof of Theorem 1. 

It is clear that JC1(Erd) and for vErd,

J(u)v=u·v+|uqx)||uq1(y)|v(y)|xy|Nαdxdy|u|p2uv.

We define, for uErd,

M(u)=|u|2+u(x)qu(y)q|xy|Nα.

Recall that

uErd=RN|u(x)|2dx+RN×RN|u(x)|q|u(y)|q|xy|Nαdxdy1/q1/2.

Define λ=M(u)1/2+αq1+2N, and let v(x)=λ2+α2q2u(λx). It is easy to see that M(v)=λ2+αq1+2NM(u)=1, and by this fact we have vErd2.

Step 1: Coercivity and Weak Lower Semicontinuity of  J(u)

By Lemma 1, we have the compact embedding, ErdRNLpRN, that is to say, there exists uniform constant C>0 such that |v|p1pvEC. Since |u|p=λN2+α2q2p|v|p, we then have |u|pCpλN2+α2q2p. Hence, we obtain

J(u)12qM(u)1p|u|p12qM(u)CppM(u)t,

where t=q1NqNα2qN2+α2q2p.

Note that by our choice of the exponents p and q, t=q1NqNα2qN2+α2q2p<1. Since for M(u) large enough, the leading term is 12qM(u) in the expression above, J is coercive; since the gradient energy is convex and the positivity of the kernel and pointwise convergence (up to a subsequence) allow applying Fatou’s lemma to the nonlocal term, and the Lp term is well controlled under the weak convergence in Erd, we conclude that J is weakly lower semicontinuous on Erd.

Step 2: Minimum and Positive Solution

We now claim that minJ(u)<0 for p<p¯=2(2q+α)2+α. Fix uErd and define the scaling v(x)=λ2+α2q2u(λx), Then, a direct calculation shows that

Jvλ=12λ2+αq1+2N|u|2+12qλ2+αq1qαN|ux)|q|u(y)|q|xy|Nαdxdy1pλ2+α2q2pN|u|p.

As we know, 2+αq1+2N=2+αq1qαN>2+α2q2pN, then for λ small, J(vλ) takes negative values. Hence, minuErdJ(u)<0.

Let unErd be a minimizing sequence such that

Junμasn.

By the above lower bound, un is bounded in H1RN. Since the embedding Hr1RNLpRN (and similarly into L2NqN+αRN) is compact for radial functions (for p<2* and under our assumptions on q), there exists a subsequence (still denoted un) and a function uHr1RN such that:

unuweaklyinH1RN,

unu strongly in LpRN and in L2NqN+αRN. And by the Hardy–Littlewood–Sobolev inequality (4), we know that the nonlocal term

RN×RN|u(x)|q|u(y)|q|xy|Nαdxdylim infn+RN×RN|un(x)|q|un(y)|q|xy|Nαdxdy.

Hence, by the weak lower semicontinuity of the norm and the functional J,

J(u)lim infnJun=μ.

Thus, u is a minimizer of J.

Since J(u)=J(|u|), without loss of generality that u0 almost everywhere. By the strong maximum principle, we know that the minimizer is a positive radial solution in Erd to the problem (4). □

Proof of Theorem 2. 

For pprd,2* and q<N+αN2, the applying Hardy–Littlewood–Sobolev inequality (6) to the nonlocal term, we have

|u(x)|q|u(y)|q|xy|NαdxdyCuL2NqN+α2qCuHr1

Since q<N+αN2, we have 2NqN+α<2*. For pprd,2*, using the Gagliardo–Nirenberg inequality yields

uLpCuL2θuL21θ,θ=N2121p<1.

Moreover, by combining this inequality with Young’s inequality, any ϵ>0 there exists C(ϵ)>0 such that

uLppϵuL22+C(ϵ)uL2s

where s=2p(1θ)2pθ. Thereby, we obtain

E(u)12uHr12ϵpuHr12C(ϵ)puHr1s.

Choosing ϵ>0 small enough such that 12ϵp>0, we conclude that E(u)+ as uHr1 and infE(u)>.

Here, we construct the (PS)c sequence (un) at level c. Suppose, for contradiction, that E(un)0 for every un with E(un)=c. Applying the deformation lemma [34], we have a (PS)c sequence (un) exists with

(7)Eunc,Eun0.

Before constructing the (PS)c sequence at level c, we apply the mountain-pass lemma, which guarantees the existence of a critical point at level c. Define the set of all continuous maps γ:[0,1]Hr1RN as

Γ=γC[0,1],Hr1RN:γ(0)=u0,γ(1)=u1,

where u0 and u1 are any two distinct points in Hr1RN such that the energy functional E(u) satisfies Eu0,Eu1<0. By the mountain-pass geometry (see Lemma 3 and Figure 1), we have

c:=infγΓmaxt[0,1]E(γ(t))maxEu0,Eu1>0.

Thus, the level c represents the mountain-pass value for the functional E, and we know that c>0.

We now show that E satisfies the (PS)c condition. Given inf EHr1> and (7), the sequence (un)Hr1RN is bounded in Hr1. There exists a subsequence (still denoted by un) and some uHr1RN such that

unuinHr1RN.

Now, we prove that unu in Hr1. Due to the radial symmetry, the embedding Hr1RNLpRN is compact, for p(2,2*). Therefore,

unustronglyinLpRN.

What left is to prove the convergence of RN×RNu(x)|q|u(y)|q|xy|Nα. First of all, according to the compact radial embedding Hr1RNL2NN+αRN, we have

unq|u|qinL2NN+αRN,

it implies

(8)|un(x)|q|u(x)|qL2NN+α0.

And by uniform boundedness in Hr1, the sequence unqL2NN+α is also bounded, i.e.,

supnunqL2NN+α<.

Now, we claim

(9)un(x)qun(y)q|xy|Nαdxdyu(x)qu(y)q|xy|Nαdxdy.

Indeed, since

un(x)qun(y)q|xy|Nαdxdyu(x)qu(y)q|xy|Nαdxdy=un(x)qun(y)qu(x)qu(y)q|xy|Nαdxdy=un(x)qu(x)qun(y)q|xy|Nαdxdy+[u(x)q]un(y)qu(y)q|xy|Nαdxdy.

Applying the Hardy–Littlewood–Sobolev inequalities (6) and (8) to the first term yields

|un(x)|q|u(x)|q|un(y)|q|xy|NαdxdyC|un(x)|q|u(x)|qL2NN+α|un(x)|qL2NN+α0.

A symmetric argument handles the second term

|u(x)|q|un(y)|q|u(y)|q|xy|Nαdxdy0.

For any test function φHr1, we have

Eun(φ)0.

Taking φ=unu, we have

Eununu=o(1)

Above all, a direct calculation shows that

Eununu=RNun·unu+ununudx+λRN×RNun(x)q2un(x)un(x)u(x)un(y)q|xy|NαdxdyRNunp2ununudx=o(1).

Then, we conclude that

unuHr10.

Since Eunc and unu strongly in Hr1, continuity of E implies E(u)=c. Further, Eun 0 and EC1Hr1 ensure E(u)=0. Hence, u is a critical point of E at level c, satisfying the (PS)c condition. □

To illustrate the qualitative conclusions of Theorem 2, we have computed a representative radial ground-state solution of the stationary Schrödinger–Poisson–Slater Equation (2). Figure 2 displays the numerical profile u(r) obtained with an imaginary-time gradient-flow scheme on the radial domain 0r20 (step size Δr=0.05) for the parameter set N=3,α=1,q=2,p=4,λ=1. The curve is smooth at the origin, satisfies the regularity condition u(0)=0, and decays monotonically to zero as r, confirming the finite-mass property. This behavior is fully consistent with the coercivity, weak lower semicontinuity, and Palais–Smale compactness established in Theorem 2, and thus provides concrete numerical evidence for the theoretical results.

Proof of Theorem 3. 

For uEα,qL2, the Pohozaev identity and subcriticality of p,q ensure E(u) is bounded below. Defining the scaling us=sN2u(sx), we have

(10)Eus=12s2|u|2+12|u|2+λ2qsNqNαIα|u|q|u|qdx1psNp/2N|u|p.

It is standard to prove that

(11)P(u)=ddss=1E(us)=|u|2+λ2q(NqNα)Iα|u|q|u|qdx1p(Np/2N)|u|p=0.

By using the Nehari identity,

(12)E(u)u=|u|2+|u|2+λIα|u|q|u|qdx|u|p=0.

and combining it with (11), we obtain

(13)1N(p2)2p|u|2dxN(p2)2p|u|2dx+NqNα2qN(p2)2pλIα|u|q|u|qdx=0.

Using the Hardy–Littlewood–Sobolev inequality and raising the Gagliardo–Nirenberg inequality to the power 2q yields

(14)λ2q(NqNα)λpNp2NCuL2r+uL2t1pNp2NuL22+1pNp2N1uL221pNp2N1(uL22+uL22)

where t=NqNα, and r=N+αq(N2). Under the condition q<N+α2Np and p>2NN2, it implies that the above inequality forces the right-hand side to be strictly positive unless uL2=uL2=0. Hence, the only solution to the Pohozaev identity (13) is u0. □

4. The Existence of the Solution in the Nonradial Case

In this section, we will prove the existence and orbital stability of ground-state solutions to the nonlocal Schrödinger–Poisson–Slater equation. The proof proceeds in two main steps: (i) establishing coercivity of the energy functional, which ensures boundedness of solutions, and (ii) demonstrating existence through minimization on the Nehari manifold.

Before proceeding with the proof of Theorem 4, we provide a visual representation of the symmetry-breaking phenomenon and the relevant parameter space.

Figure 3 visually highlights the transition from a radially symmetric solution to a nonradially symmetric one, illustrating the core idea of symmetry-breaking that will be formalized in Theorem 4.

Figure 4 reveals the critical regions for the existence and stability of ground-state solutions in the fractional nonlocal Schrödinger–Poisson–Slater equation. Specifically, the green dashed line marks the critical value for q beyond which solutions become unstable, while the orange and red lines delineate the bounds for the exponent p, indicating the regions where different types of solutions exist. The area between the orange and red lines (shown in yellow) is where nonradial solutions are stable, according to Theorem 4.

Proof of Theorem 4. 

First, we assume that the equation conserves both the mass

M(u(t))=RN|u(x,t)|2dx=M(u0)=m0,

and the energy

E(u(t))=12RN|u|2+|u|2dx+λ2qRN×RN|u(x)|q|u(y)|q|xy|Nαdxdy1pRN|u|pdx.

We consider the minimization problem on the set

Σm0=uEα,qH1RN:M(u)=m0.

Step 1:  Coercivity of the Energy Functional

Now, we claim that, under our assumptions on p and q (5), the energy E(u) is coercive. Indeed, on the one hand, by the Hardy–Littlewood–Sobolev inequality, there exists a constant C^>0 such that

λ2qRNIα|u|q|u|qdxλC^2qCuL22qθ1.

where θ1=N(q2)+α2q and 2αN<q<2NαN2. On the other hand, For the local term, by the Gagliardo–Nirenberg inequality, there exists a constant Cp>0 and an exponent θ2(0,1) such that

uLpCpuL2θ2uL21θ2

Hence,

RN|u|pdxCpuLppCpuL2pθ2

where θ2=N(p2)2p and p(2,2*). Above all, under our assumptions that 4N2α+4q2N+2α<p<2NN2, and 2αN<q<N+αN2, we obtain

(15)E(u)δuH12C,

for some δ>0 and constant C0. Thus, the energy E(u) is coercive and, being conserved, it controls the H1 norm of the solution.

Step 2: Existence of a Ground State via the Nehari Manifold

We now prove the existence of a ground-state solution by minimizing the energy functional over the Nehari manifold. The Nehari manifold is defined as

N:=uEα,qH1RN:u0,N(u)=0

and we claim the following two properties:

Claim 1. 

The infimum of the energy functional over N is strictly positive, i.e., d>0. d>0.

For uN,

| u | 2 + u 2 + λ | u ( x ) | q | u ( y ) | q | x y | N α = | u | p u H 1 p .

It implies that

0 < c 1 u H 1 2 ε 1 u H 1 p i . e . u H 1 c 1 p 2 .

Then, we obtain

E ( u ) = 1 2 1 p | u | 2 + u 2 + λ 2 q λ p | u ( x ) | q | u ( y ) | q | x y | N α 1 2 1 p c 2 p 2 > 0 .

Hence

d = inf u N E ( u ) > 0 .

Claim 2. 

There exists a function uN such that d=E(u), meaning that u minimizes the energy functional. There exists uN, such that d=E(u).

Take a minimizing sequence (uj)N, Eujd. By this way, we assume that ujH1c for some uniform constant c>0,

u n u weakly in H 1 R N ,

u n u a . e . in R N ,

u n u in L loc p R N for p < 2 N N 2 .

For the nonlocal term, the Hardy–Littlewood–Sobolev inequality yields

R N × R N | u ( x ) | q | u ( y ) | q | x y | N α d x d y C u L s 2 q , s = 2 N q N + α .

Since s<2NN2 when q<N+αN2, applying the Sobolev embedding H1(RN)Ls(RN), we obtain

u L s C u H 1 ,

and so

| u | q | u | q | x y | N α C u H 1 2 q .

By the variational characterization of λ0,

λ 0 = inf u H 1 ( R N ) { 0 } 2 q p | u | p | u | q | u | q | x y | N α ,

for λλ0, we obtain

λ | u | q | u | q | x y | N α u L p p u H 1 2 .

Hence, for λλ0, the energy E(u) is nonnegative.

According to the nonlocal Brezis–Lieb lemma [19], we know that

lim infn|un|q|un|q|xy|Nα|u|q|u|q|xy|Nα+lim infn|unu|q|unu|q|xy|Nα.

By Proposition 4.3 and Proposition 3.3 in [19], equality in the nonlocal Brezis–Lieb inequality holds if unu in LlocpRN for p>q. Here, by Rellich–Kondrachov theorem, it is satisfied when 1p>121α+. Thus, we conclude that

limn|unu|q|unu|q|xy|Nα=0.

Using the Sobolev embedding H1(RN)Ls(RN) again, we have

unustronglyinH1(RN).

Then, we deduce

N(u)lim inf|un|2+un2+λ|un(x)|q|un(y)|q|xy|Nα|un|p=0.

By contradiction, if N(u)<0, and

N(tu)=t2|u|2+ε2u2+λt2q|u(x)|q|u(y)|q|xy|Ndtp|u|p>0,

for t>0 small, then we have N(tcu)=0 for some 0<t<1. Then, tcuN and Etcud. But we compute:

dEtcu=121ptc2uH12+λ2qλptcp|u|p<E(u)=d,

which is absurd. Then, N(u)=0, E(u)=d, i.e., u is a minimizer of E on N. Following the Lagrange multiplier, E(u)v=λ˜N(u)v, and taking u=v,

N(u)=E(u)u=λ˜N(u)u=0.

that is

λ˜2|u|2+u2+2qλ|u(x)|q|u(y)|q|xy|Nαp|u|p=0.

It implies that

λ˜(2p)uH12+λ(2qp)|u(x)|q|u(y)|q|xy|Nα=0.

As we know p>2, 2q>p and λ>0, u0, we obtain λ˜=0. Therefore, we obtain E(u)u=λ˜N(u)u=0 for all vH1RN. Hence, u is a critical point. Then, ϕ is a ground state.

Step 3: Existence of Solutions to the Cauchy Problem

Consider the Cauchy problem

itu=Δu+u+λIα|u|q|u|q2u|u|p2u,(x,t)RN×R,u(0)=u0H1(RN).

Define the linear operator

L=Δ+ε2.

Write the equation in Duhamel form, we have

u(t)=eitLu0i0tei(ts)LS(u(s))ds,

where

S(u)=λIα|u|q|u|q2u|u|p2u.

Fix T>0 and define

XT:=C[0,T];H1RNST

where

ST=Lq00,T;W1,r0RN

and we choose the Schrödinger-admissible pair (q0,r0)

r0=2Nq2Nαandq0=4qNq2N+α.

It follows from the classical Strichartz estimates, that is

eitLu0XTCu0H1.

In order to balance the interplay between the nonlocal term and the local term, define NT=Lq˜(0,T;W1,r˜(RN)) with dual exponents q˜=q02q1 and r˜=r02q1. By the Hardy–Littlewood–Sobolev inequality and Sobolev’s embedding, we obtain

λIα|u|q|u|q2uNTCuXT2q1,

and using Sobolev’s embedding,

|u|p2uNTCuXTp1.

hence, the Duhamel operator is bounded. This implies that

uIα|u|q|u|q2u

is locally Lipschitz from XT into the dual space NT.

S(u)NTCRuXT2q1,

where CR. Thus, the overall estimate for the inhomogeneous term becomes

0tei(ts)LS(u(s))dsSTCRuXT2q1,

where Tη is time integration.

Above all, we obtain

Φ(u)XTCu0H1+CTηCRR2q1.

here η=12q1q0=1(2q1)(Nq2N+α)4q. For any u,vBR,

Φ(u)(t)Φ(v)(t)=i0tei(ts)L(S(u(s))S(v(s)))ds

so

Φ(u)Φ(v)XTCS(u)S(v)NTCTηuXT2q2+vXT2q2+uXTp2+vXTp2uvXTCTη4R2q2uvXT.

so the mapping

Φ(u)(t)=eitLu0i0tei(ts)LS(u(s))ds

is a contraction on a ball BRXT, where

BR=uXT:uXTR,

for some R>0 provided T>0 is sufficiently small. Thus, there exists a unique local solution uXT. It implies that if the initial data have finite energy Then, the solution cannot blow up in finite time.

By the conservation of energy and mass, and the coercivity inequality (15),

u(t)H12Eu0+Cδ,t,

the local solution extends globally in time, i.e.,

uCR;H1(RN).

Moreover, the solution depends continuously on the initial data.

Step 4: Orbital Stability of the Ground State

In the end, we prove orbital stability of the ground state. By contradiction, assume that orbital stability fails. Then, there exists an ϵ0>0, a sequence of initial data un0H1RN, and a sequence of positive times tn such that

un0ϕH10asn

but the corresponding solutions un(t) of Equation (2) with initial data un0 satisfy

infθRuntneiθϕH1>ϵ0foralln

Denote ϕn:=untn. At time tn, we have

Eϕnmm0andMϕnM(Q).

Next, denote by un0 the modulated initial data sequence. Since ϕ is a ground state to (2), by definition, we have

E(ϕ)=m(m0):=infu0uH1E(u)andϕH1:=b>0.

By Sobolev embedding, we still have

un0QLppun0QH1p0

so un0Q in Lp. Then, conservation laws imply that, up to translations and phase shifts,

ϕnQH10.

In order to conclude strong convergence of the sequence, we must rule out vanishing and dichotomy in the concentration-compactness principle.

Rule out Vanishing. If the mass of ϕn “spreads out” to infinity, then Eϕn would not converge to m(m0) contradicting energy conservation.

Assume vanishing, which means

limnsupyRNB(y,R)ϕn(x)2dx=0R>0.

By Lions’ vanishing lemma [35], for 2<p<2*, limnϕnLp=0. The nonlocal term is controlled by the Hardy–Littlewood–Sobolev inequality (6):

ϕn(x)qϕn(y)q|xy|NαdxdyCϕnLs2q,s=2NqN+α.

Since s<2* (from q<N+α2Np), vanishing implies ϕnLs0, so

ϕn(x)qϕn(y)q|xy|Nαdxdy0.

Rule out Dichotomy. Let m(m0) denote the minimal energy among functions with fixed mass m0, i.e.,

m(m0)=inf{E(u):uH1(RN),M(u)=m0},

where

E(u)=12RN|u|2+|u|2dx+λ2qRN×RN|u(x)|q|u(y)|q|xy|Nαdxdy1pRN|u|pdx,

and the mass is

M(u)=RN|u|2dx.

Assume by contradiction that there exists a minimizing sequence {ϕn} for m(m0), which exhibits dichotomy. That is, after extracting a subsequence, there exist sequences vn and wn with asymptotically disjoint supports such that

ϕn(vn+wn)H10,vnL22α,wnL22m0α,

for some 0<α<m0. In this scenario, the splitting of ϕn into two disjoint parts implies that the strict subadditivity of mλ is violated. Denote

A(u)=12RN|u|2+|u|2dx;B(u)=λ2qRN×RN|u(x)|q|u(y)|q|xy|Nαdxdy;C(u)=1pRN|u|pdx;

Define the scaling

uθ(x)=θauθbx;θ(1,+),

where a=12+N22b=1+b(N2)2 and b=1qq(N2)(N+α). Through direct computation, we obtain

Euθ=θ2a+2bbNA(u)+θ2aqb(N+α)B(u)θapbNC(u)=θA(u)+θB(u)θsC(u).

Under the parameter assumption, we deduce that

s=p21(2N)(q1)q(2N)+(3Nα)N(q1)q(2N)+(3Nα)>1.

By the definition of m(m0) and taking the infimum over all uΣm0 yields

m(θα)<θm(α).

Here, we employ a key lemma in concentration-compactness from [12]. The lemma asserts that if a function h:[0,m0]R satisfies

h(θα)<θh(α)forallα(0,m0)andθ1,m0α,

then it is strictly subadditive

h(m0)<h(α)+h(m0α)forallα(0,m0).

The strict subadditivity assumption then reads

m(m0)<m(α)+m(m0α)forallα(0,m0).

Since {ϕn} is a minimizing sequence with M(ϕn)m0 and E(ϕn)m(m0), one expects that if dichotomy occurs Then, the energy asymptotically splits as

E(ϕn)=E(vn)+E(wn)+o(1).

By the definition of minimal energy, we have

E(vn)mvnL22andE(wn)mwnL22.

Passing to the limit, we obtain

lim infnE(ϕn)m(α)+m(m0α),

which contradicts the dichotomy.

Since neither vanishing nor dichotomy occurs, the concentration-compactness principle ensures that there exists a sequence ynRN such that the translated sequence

vn(x)=ϕnx+ynϕstronglyinH1(RN).

Thus, ϕ achieves the minimal energy mm0 and is a ground state. Consequently, for large n,

infθR,yRNuntneiθQ(·y)H10.

Thus, by ruling out vanishing and dichotomy and applying the concentration compactness principle, we conclude that the sequence ϕn converges strongly in H1RN, and the ground-state solution is orbitally stable.

Author Contributions

F.D.: Conceptualization, Methodology, Formal analysis, Validation, Writing—original draft; Z.W.: Conceptualization, Methodology, Writing—review & editing; H.L.: Writing—review & editing; L.C.: Conceptualization, Methodology, Writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Data Availability Statement

No data was used for the research described in the article.

Conflicts of Interest

The authors declare no conflicts of interest.

Footnotes

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Figures and Table

Figure 1 Three-dimensional (3D) energy functional E(u) landscape, illustrating the mountain pass geometry with critical points marked as red ‘×’, highlighting the transition between local minima and critical points.

View Image -

Figure 2 Radial ground-state profile u(r) for the stationary Schrödinger–Poisson–Slater Equation (2) with N=3, α=1, q=2, p=4 and λ=1. The solution is obtained via an imaginary-time gradient-flow scheme on a radial mesh r[0,20] (Δr=0.05). The curve is smooth at the origin (u(0)=0) and decays monotonically to zero.

View Image -

Figure 3 Symmetry-breaking comparison of ground states. (Left): Radial energy-density contour for N=3, α=1.5, p=3.0, q=1.5, λ=0.5. The red dot marks the minimizer in the radial subspace (perfectly circular symmetry). (Right): Nonradial energy-density contour under the same parameters, where the red dot indicates the global minimizer in the full space—evidently off the circle, illustrating symmetry-breaking.

View Image -

Figure 4 Parameter space for the existence and stability of ground-state solutions in the nonlocal Schrödinger–Poisson–Slater equation (Theorem 4) for N=3 and α=1.

View Image -

Simplified parameter ranges and conclusions of main theorems.

Theorem Parameter Range Conclusion
Existence and Stability (p(prd,p¯),q<N+αN2) or (p(2*,p¯),q>N+αN2) J is coercive, weakly lower semicontinuous, negative minimum exists
Weak Lower Semicontinuity and Coercivity p ( p r d , 2 * ) , q < N + α N 2 E(u) is coercive, weakly lower semicontinuous, minimum exists, Palais–Smale condition holds
Nonexistence (Pohozaev Identity) p > 2 N N 2 , q < N + α 2 N p No nontrivial ground state (u0)
Symmetry-Breaking and Stability q < N + α N 2 , 2 ( 2 q + 2 N α ) 2 N + 2 α < p < 2 N N 2 Existence of orbitally stable ground state

References

1. Adams, D.R.; Hedberg, L.I. Function Spaces and Potential Theory; Springer: Berlin/Heidelberg, Germany, 1999; Volume 314.

2. Moroz, V.; Van Schaftingen, J. Groundstates of nonlinear Choquard equations: Existence, qualitative properties and decay asymptotics. J. Funct. Anal.; 2013; 265, pp. 153-184. [DOI: https://dx.doi.org/10.1016/j.jfa.2013.04.007]

3. Di Nezza, E.; Palatucci, G.; Valdinoci, E. Hitchhiker’s guide to the fractional Sobolev spaces. Bull. Sci. Math.; 2012; 136, pp. 521-573. [DOI: https://dx.doi.org/10.1016/j.bulsci.2011.12.004]

4. Slater, J.C. A simplification of the Hartree-Fock method. Phys. Rev.; 1951; 81, 385. [DOI: https://dx.doi.org/10.1103/PhysRev.81.385]

5. Bokanowski, O.; López, J.L.; Soler, J. On an exchange interaction model for quantum transport: The Schrödinger–Poisson–Slater system. Math. Model. Methods Appl. Sci.; 2003; 13, pp. 1397-1412. [DOI: https://dx.doi.org/10.1142/S0218202503002969]

6. Mauser, N.J. The Schrödinger-Poisson-Xα equation. Appl. Math. Lett.; 2001; 14, pp. 759-763. [DOI: https://dx.doi.org/10.1016/S0893-9659(01)80038-0]

7. Benguria, R.; Brézis, H.; Lieb, E.H. The Thomas-Fermi-von Weizsäcker theory of atoms and molecules. Commun. Math. Phys.; 1981; 79, pp. 167-180. [DOI: https://dx.doi.org/10.1007/BF01942059]

8. Lieb, E.H. Thomas-Fermi and related theories of atoms and molecules. Rev. Mod. Phys.; 1981; 53, 603. [DOI: https://dx.doi.org/10.1103/RevModPhys.53.603]

9. Eberhard, E.; Dreizler, R. Density Functional Theory: An Advanced Course; Springer: Berlin/Heidelberg, Germany, 2011.

10. Nier, F. A variational formulation of Schrödinger-Poisson systems in dimension d ≤ 3. Commun. Partial. Differ. Equ.; 1993; 18, pp. 1125-1147. [DOI: https://dx.doi.org/10.1080/03605309308820966]

11. Ruiz, D. The Schrödinger–Poisson equation under the effect of a nonlinear local term. J. Funct. Anal.; 2006; 237, pp. 655-674. [DOI: https://dx.doi.org/10.1016/j.jfa.2006.04.005]

12. Lions, P.L. The concentration-compactness principle in the calculus of variations. The limit case, part 1. Rev. Mat. Iberoam.; 1985; 1, pp. 145-201. [DOI: https://dx.doi.org/10.4171/rmi/6]

13. Lieb, E.H.; Loss, M. Analysis; American Mathematical Society: Providence, RI, USA, 2001; Volume 14.

14. Lieb, E.H.; Simon, B. The Thomas-Fermi theory of atoms, molecules and solids. Adv. Math.; 1977; 23, pp. 22-116. [DOI: https://dx.doi.org/10.1016/0001-8708(77)90108-6]

15. Lions, P. The Choquard equation and related questions. Nonlinear Anal. Theory Methods Appl.; 1980; 4, pp. 1063-1072. [DOI: https://dx.doi.org/10.1016/0362-546X(80)90016-4]

16. Jeanjean, L.; Le, T.T. Multiple normalized solutions for a Sobolev critical Schrödinger-Poisson-Slater equation. J. Differ. Equations; 2021; 303, pp. 277-325. [DOI: https://dx.doi.org/10.1016/j.jde.2021.09.022]

17. Berestycki, H.; Lions, P.L. Nonlinear scalar field equations. Pt. 1. Arch. Ration. Mech. Anal.; 1983; 82, pp. 313-346. [DOI: https://dx.doi.org/10.1007/BF00250555]

18. Ruiz, D. On the Schrödinger-Poisson-Slater system: Behavior of minimizers, radial and nonradial cases. Arch. Ration. Mech. Anal.; 2010; 198, pp. 349-368. [DOI: https://dx.doi.org/10.1007/s00205-010-0299-5]

19. Mercuri, C.; Moroz, V.; Van Schaftingen, J. Groundstates and radial solutions to nonlinear Schrödinger–Poisson–Slater equations at the critical frequency. Calc. Var. Partial. Differ. Equations; 2016; 55, pp. 1-58. [DOI: https://dx.doi.org/10.1007/s00526-016-1079-3]

20. Bellazzini, J.; Ghimenti, M.; Mercuri, C.; Moroz, V.; Van Schaftingen, J. Sharp Gagliardo–Nirenberg inequalities in fractional Coulomb–Sobolev spaces. Trans. Am. Math. Soc.; 2018; 370, pp. 8285-8310. [DOI: https://dx.doi.org/10.1090/tran/7426]

21. Servadei, R.; Valdinoci, E. Variational methods for non-local operators of elliptic type. Discret. Contin. Dyn. Syst; 2013; 33, pp. 2105-2137. [DOI: https://dx.doi.org/10.3934/dcds.2013.33.2105]

22. Tang, Y.; Huang, Y.; Liu, Z.; Moroz, V. Symmetry breaking and multiple solutions for the Schrödinger–Poisson–Slater equation. Z. Angew. Math. Phys.; 2024; 75, 114. [DOI: https://dx.doi.org/10.1007/s00033-024-02261-4]

23. Liu, Z.; Moroz, V. Asymptotic profile of ground states for the Schrödinger–Poisson–Slater equation. Nonlinear Anal.; 2022; 218, 112778. [DOI: https://dx.doi.org/10.1016/j.na.2021.112778]

24. Zhang, L.; Hou, W. Standing waves of nonlinear fractional p-Laplacian Schrödinger equation involving logarithmic nonlinearity. Appl. Math. Lett.; 2020; 102, 106149. [DOI: https://dx.doi.org/10.1016/j.aml.2019.106149]

25. Wang, G.; Ren, X.; Bai, Z.; Hou, W. Radial symmetry of standing waves for nonlinear fractional Hardy–Schrödinger equation. Appl. Math. Lett.; 2019; 96, pp. 131-137. [DOI: https://dx.doi.org/10.1016/j.aml.2019.04.024]

26. Zhang, L.; Ahmad, B.; Wang, G.; Ren, X. Radial symmetry of solution for fractional p- Laplacian system. Nonlinear Anal.; 2020; 196, 111801. [DOI: https://dx.doi.org/10.1016/j.na.2020.111801]

27. Zhang, L.; Liu, Q.; Ahmad, B.; Wang, G. Nonnegative solutions of a coupled k-Hessian system involving different fractional Laplacians. Fract. Calc. Appl. Anal.; 2024; 27, pp. 1835-1851. [DOI: https://dx.doi.org/10.1007/s13540-024-00277-1]

28. Zhang, L.; Nie, X. Hopf’s lemma and radial symmetry for the Logarithmic Laplacian problem. Fract. Calc. Appl. Anal.; 2024; 27, pp. 1906-1916. [DOI: https://dx.doi.org/10.1007/s13540-024-00285-1]

29. Wang, G.; Yang, R.; Zhang, L. The properties of a new fractional g-Laplacian Monge-Ampère operator and its applications. Adv. Nonlinear Anal.; 2024; 13, 20240031. [DOI: https://dx.doi.org/10.1515/anona-2024-0031]

30. Siciliano, G. Multiple positive solutions for a Schrödinger–Poisson–Slater system. J. Math. Anal. Appl.; 2010; 365, pp. 288-299. [DOI: https://dx.doi.org/10.1016/j.jmaa.2009.10.061]

31. Georgiev, V.; Prinari, F.; Visciglia, N. On the radiality of constrained minimizers to the Schrödinger–Poisson–Slater energy. Ann. L’Inst. Henri Poincaré Anal. Non Linéaire; 2012; 29, pp. 369-376. [DOI: https://dx.doi.org/10.1016/j.anihpc.2011.12.001]

32. Lei, C.; Lei, Y. On the existence of ground states of an equation of Schrödinger–Poisson–Slater type. Comptes Rendus. Mathématique; 2021; 359, pp. 219-227. [DOI: https://dx.doi.org/10.5802/crmath.175]

33. Lions, P.L. Solutions of Hartree-Fock equations for Coulomb systems. Commun. Math. Phys.; 1987; 109, pp. 33-97. [DOI: https://dx.doi.org/10.1007/BF01205672]

34. Willem, M. Minimax Theorems; Springer: Berlin/Heidelberg, Germany, 2012; Volume 24.

35. Lions, P.L. The concentration-compactness principle in the Calculus of Variations. The locally compact case, part 1. Ann. L’Inst. Henri Poincaré Anal. Non Linéaire; 1984; 1, pp. 109-145. [DOI: https://dx.doi.org/10.1016/s0294-1449(16)30428-0]

© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.