Citation: Pauline M. Insch, Gillian Slessor, Louise H. Phillips, Anthony Atkinson, Jill Warrington. The Impact of Aging and Alzheimers Disease on Decoding Emotion Cues from Bodily Motion[J]. AIMS Neuroscience, 2015, 2(3): 139-152. doi: 10.3934/Neuroscience.2015.3.139
[1] | Dong-Mei Li, Bing Chai . A dynamic model of hepatitis B virus with drug-resistant treatment. AIMS Mathematics, 2020, 5(5): 4734-4753. doi: 10.3934/math.2020303 |
[2] | Komal Bansal, Trilok Mathur, Narinderjit Singh Sawaran Singh, Shivi Agarwal . Analysis of illegal drug transmission model using fractional delay differential equations. AIMS Mathematics, 2022, 7(10): 18173-18193. doi: 10.3934/math.20221000 |
[3] | Kasbawati, Mariani, Nur Erawaty, Naimah Aris . A mathematical study of effects of delays arising from the interaction of anti-drug antibody and therapeutic protein in the immune response system. AIMS Mathematics, 2020, 5(6): 7191-7213. doi: 10.3934/math.2020460 |
[4] | Sara J Hamis, Fiona R Macfarlane . A single-cell mathematical model of SARS-CoV-2 induced pyroptosis and the effects of anti-inflammatory intervention. AIMS Mathematics, 2021, 6(6): 6050-6086. doi: 10.3934/math.2021356 |
[5] | Yaw Chang, Wei Feng, Michael Freeze, Xin Lu, Charles Smith . Elimination, permanence, and exclusion in a competition model under Allee effects. AIMS Mathematics, 2023, 8(4): 7787-7805. doi: 10.3934/math.2023391 |
[6] | Yasir Nadeem Anjam, Asma Arshad, Rubayyi T. Alqahtani, Muhammad Arshad . Unveiling the dynamics of drug transmission: A fractal-fractional approach integrating criminal law perspectives. AIMS Mathematics, 2024, 9(5): 13102-13128. doi: 10.3934/math.2024640 |
[7] | Hongyan Wang, Shaoping Jiang, Yudie Hu, Supaporn Lonapalawong . Analysis of drug-resistant tuberculosis in a two-patch environment using Caputo fractional-order modeling. AIMS Mathematics, 2024, 9(11): 32696-32733. doi: 10.3934/math.20241565 |
[8] | Ahmed M. Elaiw, Amani S. Alsulami, Aatef D. Hobiny . Global properties of delayed models for SARS-CoV-2 infection mediated by ACE2 receptor with humoral immunity. AIMS Mathematics, 2024, 9(1): 1046-1087. doi: 10.3934/math.2024052 |
[9] | Biwen Li, Qiaoping Huang . Synchronization issue of uncertain time-delay systems based on flexible impulsive control. AIMS Mathematics, 2024, 9(10): 26538-26556. doi: 10.3934/math.20241291 |
[10] | Mansoor H. Alshehri . Computational study on encapsulation of 5-fluorouracil drug in nanotubes. AIMS Mathematics, 2022, 7(9): 16975-16985. doi: 10.3934/math.2022932 |
This article is concerned with a class of generalized quasilinear Schrödinger equations
$ {−div(g2(u)∇u)+g(u)g′(u)|∇u|2+V(x)u=K(x)f(u)+λW(x)h(u),in RN,u∈D1,2(RN), $
|
(1.1) |
where $ N\geq3 $, $ \lambda > 0 $, $ f, h $: $ \mathbb{R}\rightarrow \mathbb{R} $ and $ V, K, W $: $ \mathbb{R}^{N}\rightarrow \mathbb{R} $ are nonnegative continuous and $ g(s)\in C^1(\mathbb{R}, \mathbb{R}^+) $, which is nondecreasing with respect to $ |s| $.
These equations are related to the existence of solitary waves for the Schrödinger equation
$ i∂tz=−Δz+V(x)z−k(x,z)z−Δl(|z|2)l′(|z|2)z, x∈RN, $
|
(1.2) |
where $ z $: $ \mathbb{R}\times\mathbb{R}^N\rightarrow \mathbb{C} $, $ V $: $ \mathbb{R}^N\rightarrow \mathbb{R} $ is a given potential, $ l $: $ \mathbb{R}\rightarrow \mathbb{R} $ and $ k $: $ \mathbb{R}^N\times\mathbb{C}\rightarrow \mathbb{R} $ are fixed functions. Quasilinear equations of the form (1.2) appear naturally in mathematical physics and have been derived as models of several physical phenomena corresponding to various types of $ l $. For instance, the case $ l(s) = s $ appears in the superfluid film equation in plasma physics [18]. If $ l(s) = \sqrt{1+s} $, the equation models the propagation of a high-irradiance laser in a plasma, as well as the self-channeling of a high-power ultrashort laser in matter[19]. For more physical motivations and more references dealing with various applications, we refer to [5,16,17,26,28].
If we set $ z(t, x) = e^{-iEt}u(x) $ in (1.2), we obtain the corresponding equation of elliptic type
$ −Δu+V(x)u−Δ(l(u2))l′(u2)u=k(x,u)u, x∈RN. $
|
(1.3) |
Notice that if we let
$ g2(u)=1+[(l(u)2)′]22, $
|
we have the following equation
$ −div(g2(u)∇u)+g(u)g′(u)|∇u|2+V(x)u=k(x,u)u. $
|
(1.4) |
One of the most interesting cases is that $ g(s) = \sqrt{1+2s^2} $, and then (1.4) changes to
$ −Δu+V(x)u−[Δ(u)2]u=k(x,u)u. $
|
(1.5) |
The Schrödinger equation is quasilinear as the term $ [\Delta(u)^2]u $ is linear about the second derivatives. Over the past decades, many interesting results about the existence of solutions to (1.5) have been established. It is difficult to give a complete reference, so we only refer to some early works [23,24] for special $ k(x, u)u $ and some papers [1,6,9,13,22,35] closely related to our paper. Particularly, Wang and Yao [36] studied the existence of nontrivial solutions to (1.5) with concave-convex nonlinearities $ \mu |u|^{\hat{p}-2}u+|u|^{\hat{q}-2}u $, $ 2 < \hat{p} < 4 $, $ 4 < \hat{q} < 22^* $, and the potential $ V(x) $ satisfied the following conditions:
$ (V'_1) $ $ V\in C(\mathbb{R}^N, \mathbb{R}) $ and $ 0 < V_0\leq\inf\limits_{x\in\mathbb{R}^N}V(x) $;
$ (V'_2) $ There exists $ V_1 > 0 $ such that $ V(x) = V(|x|)\leq V_1 $ for all $ x\in\mathbb{R}^N $;
$ (V'_3) $ $ \nabla V(x)x\leq 0 $ for all $ x\in\mathbb{R}^N $.
In this paper we investigate the more general Eq (1.4) where the nonlinearity is like $ \mu W(x)|u|^{\hat{p}-2}u+K(x)|u|^{\hat{q}-2}u $, $ 1 < \hat{p} < 2 $, $ 4 < \hat{q} < 22^* $ and $ V, K, W $: $ \mathbb{R}^{N}\rightarrow \mathbb{R} $ satisfy some conditions listed below. There are also many works on the equation in the recent years, but we only mention those closely related to our paper, [7,8,10,29,30] and the references therein. Particularly, Furtado et al. [14] investigated solutions to (1.4) with a huge class of functions $ g $ satisfying the following condition $ (g_0) $.
$ (g_0) $ $ g\in C^1(\mathbb{R}, (0, +\infty)) $ is even, non-decreasing in $ [0, +\infty) $, $ g(0) = 1 $ and satisfies
$ g∞:=limt→∞g(t)t∈(0,∞) $
|
(1.6) |
and
$ β:=supt∈Rtg′(t)g(t)≤1. $
|
(1.7) |
When $ g $ satisfies $ (g_{0}) $, the existence of solutions to (1.4) has been investigated by several authors over the past years [15,27] and the references therein. In particular, in [25] the authors considered the positive solutions to it when the nonlinearity is like $ \mu |u|^{\hat{p}-2}u+|u|^{\hat{q}-2}u $, $ 1 < \hat{p} < 2 $, $ 4 < \hat{q} < 22^* $ where the potential $ V(x) $ satisfied $ (V'_{1}) $ and the following condition:
$ (V'_4) $ $ [V(x)]^{-1}\in L^1(\mathbb{R}^N) $.
An important class of problems associated to (1.1) is the case when $ V(x) $ vanishes at infinity
$ \lim\limits_{|x|\rightarrow +\infty}V(x) = 0, $ |
which has been extensively investigated for the corresponding second order nonlinear Schrödinger equations after the researches of e.g., [2,3]. See also [11,21,32,33,34] for some work about $ V(x) $ vanishing at infinity. However, there are only few works in this case for the more general Eq (1.1). Motivated by the above articles, we investigate the existence of solutions to (1.1) when the potential $ V $ vanishes at infinity for a huge class of $ g $ (satisfying $ (g_{0}) $).
In this paper, we consider the generalized quasilinear Schrödinger Eq (1.1) with vanishing potentials and concave-convex nonlinearity $ K(x)f(u)+\lambda W(x)h(u) $. Since the problem is set on the whole space $ \mathbb{R}^N $, we have to deal with the loss of compactness. In this respect we use the class of functions $ V, K $ introduced in [2] for second order Schrödinger equations, which is more general than those in [3].
As in [2], it is said that $ (V, K)\in\mathcal{K} $ if the following conditions hold:
(I) $ K(x), \ V(x) > 0, \ \forall x\in\mathbb{R}^N $ and $ K\in L^\infty(\mathbb{R}^N) $.
(II) If $ \{A_n\}\subset\mathbb{R}^N $ is a sequence of Borel sets, such that $ |A_n|\leq R $ for some $ R > 0 $ and for all $ n\in\mathbb{N} $, then
$ limr→+∞∫An∩Bcr(0)K(x)dx=0, uniformly in n∈N. $
|
($K_1$) |
(III) One of the below conditions satisfies:
$ KV∈L∞(RN) $
|
($K_2$) |
or there is $ \sigma\in(2, 2^*) $ such that
$ K(x)[V(x)]2∗−σ2∗−2→0, as |x|→+∞. $
|
($K_3$) |
We also use the following conditions on $ V $ and $ W $:
$ (V_1) $ $ V(x)\in L^\infty(\mathbb{R}^N) $;
$ (W_0) $ $ W(x) > 0 $ for all $ x\in\mathbb{R}^N $;
$ (W_1) $ $ W(x)\in L^{1}(\mathbb{R}^{N})\cap L^{\infty}(\mathbb{R}^{N}) $;
$ (W_2) $ $ \frac{W(x)}{V(x)}\in L^{\infty}(\mathbb{R}^{N}) $.
We impose the following conditions on $ h $ and $ f $:
$ (H_0) $ $ h\in C(\mathbb{R}, \mathbb{R}^+) $ and $ h(t) = 0 $ for all $ t\leq 0 $;
$ (H_1) $ There exists $ b_1, \ b_2 > 0 $ such that $ h(t)\leq b_1|t|^{\tau_1-1}+b_2|t|^{\tau_2-1} $, $ \tau_1, \tau_2\in (1, 2) $ for any $ t\in\mathbb{R} $;
$ (F_0) $ $ f\in C(\mathbb{R}, \mathbb{R}^+) $ and $ f(t) = 0 $ for all $ t\leq 0 $;
$ (F_1) $ $ \lim\limits_{|t|\rightarrow +\infty}\frac{f(t)}{|t|^{22^*-1}} = 0 $;
$ (F_2) $ $ \lim\limits_{|t|\rightarrow 0}\frac{f(t)}{|t|} = 0 $ if ($K_2$) holds or $ \lim\limits_{|t|\rightarrow 0}\frac{f(t)}{|t|^{\sigma-1}} = 0 $ if ($K_3$) holds;
$ (F_3) $ $ \frac{F(t)}{t^{4}}\rightarrow +\infty $, as $ t\rightarrow +\infty $;
$ (F_4) $ There exists $ \mu > 2+2\beta $ such that $ \frac{1}{\mu}f(t)t\geq F(t) $, where $ \beta $ is in (1.7).
Observe that there are many natural functions $ f(t), h(t) $ satisfying the above conditions. For example, $ f(t) = |t|^{2^*+1} $ and $ h(t) = |t|^{\frac{1}{2}} $ may serve as examples satisfying $ (F_{1}) $–$ (F_{4}) $ and $ (H_{1}) $, respectively.
Our main theorem is stated as follows.
Theorem 1.1. Assume that $ (V, K)\in\mathcal{K}, $ $ (g_{0}), $ $ (V_{1}), $ $ (W_{0}) $–$ (W_{2}), $ $ (F_{0}) $–$ (F_{4}), $ $ (H_{0}) $ and $ (H_{1}) $ hold. Then, there exists $ \lambda_0 > 0 $ such that (1.1) possesses a positive solution for any $ \lambda\in(0, \lambda_0) $.
Furthermore, for the case where ($K_2$) holds, we can prove that (1.1) possesses a ground state solution. To this end, we assume the following conditions on $ h $ and $ f $:
$ (H_0') $ $ h\in C(\mathbb{R}, \mathbb{R}) $, $ h(t) $ is odd and $ h(t)\geq0 $ for all $ t\geq 0 $.
$ (H_1') $ There exists $ b_3 > 0 $ and $ \tau_3\in(1, 2) $ such that $ h(t)\leq b_3|t|^{\tau_3-1} $.
$ (H_2') $ There exists a constant $ \tilde{C} > 0 $ such that $ \lim\limits_{t\rightarrow0}\frac{H(t)}{|t|^{\tau_3}} = \tilde{C} $.
$ (F_0') $ $ f\in C(\mathbb{R}, \mathbb{R}) $, $ f(t) $ is odd and $ f(t)\geq0 $ for all $ t\geq 0 $.
Proposition 1.2. Assume that $ (V, K)\in\mathcal{K} $ where ($K_2$) holds and $ (g_{0}), $ $ (V_{1}), $ $ (W_{0}) $–$ (W_{2}), $ $ (F'_{0}), $ $ (F_{1}), $ $ (F_{2}), $ $ (F_{4}), $ $ (H'_{0}) $–$ (H'_{2}) $ hold. Then, there exists $ \lambda_1 > 0 $ such that (1.1) possesses a ground state solution for any $ \lambda\in(0, \lambda_1) $.
We emphasize that the main result in this paper is essentially different from the aforementioned works. Indeed, in [25,36] the authors considered two kinds of quasilinear Schrödinger equations with concave-convex nonlinearities, but required that the potential $ V(x) $ have a positive lower bound. In [11,21] the authors showed the existence of nontrivial solutions for different problems with vanishing potentials. In this paper, we investigate a different class of generalized quasilinear Schrödinger equations with vanishing potentials and concave-convex nonlinearities. As far as we know, few works in this case seem to have appeared in the literature.
The paper is organized as follows. In Section 2, some preliminary results are presented. In Section 3, we verify that the functional associated to the problem satisfies the geometric conditions of the mountain pass theorem, and the boundedness of the Cerami sequences associated with the corresponding minimax level is proved. Lastly, in Section 4, the existence of a positive solution and a ground state solution for (1.1) is established.
As usual, we use the Sobolev space
$ X={u∈D1,2(RN):∫RNV(x)u2dx<+∞} $
|
(2.1) |
endowed with the norm
$ ‖u‖=(∫RN(|∇u|2+V(x)|u|2)dx)12. $
|
(2.2) |
The weighted Lebesgue space is defined as follows
$ L_K^q(\mathbb{R}^N) = \big\{u:\mathbb{R}^N\rightarrow \mathbb{R}\ |\ u\ \text{is measurable and }\int_{\mathbb{R}^N}K(x)|u|^q dx < +\infty\big\} $ |
endowed with the norm
$ \|u\|_{K, q}: = \big(\int_{\mathbb{R}^N}K(x)|u|^q dx\big)^\frac{1}{q}. $ |
The space $ L_W^p(\mathbb{R}^N) $ with the norm $ \|u\|_{W, p} $ is similarly defined.
The following proposition is proved in [2].
Proposition 2.1. [2] Assume that $ (V, K)\in\mathcal{K} $. Then, $ X $ is compactly embedded in $ L_K^q(\mathbb{R}^N) $ for all $ q\in(2, 2^*) $ if ($K_2$) holds. If ($K_3$) holds, $ X $ is compactly embedded in $ L_K^\sigma(\mathbb{R}^N) $.
To resolve (1.1), due to the appearance of the nonlocal term $ \int_{\mathbb{R}^N}g^2(u)|\nabla u|^2dx $, the right working space seems to be
$ X0={u∈X:∫RNg2(u)|∇u|2dx<∞}. $
|
However, generally $ X_0 $ is not a linear space and the functional
$ Iλ(u)=12∫g(u)2|∇u|2dx+12∫V(x)u2dx−∫K(x)F(u)dx−λ∫W(x)H(u)dx $
|
(2.3) |
may be not well defined on $ X_0 $, where
$ F(u) = \int_0^u f(s)ds, \ \ H(u) = \int_0^u h(s)ds. $ |
To avoid these drawbacks, following [20,26,30], we make a change of variables
$ v = G(u) = \int_0^ug(t)dt. $ |
Then, it follows from the properties of $ g $, $ G $ and $ G^{-1} $, which will be listed in Lemma 2.4 that if $ v\in X $, then $ u = G^{-1}(v)\in X $ and
$ \int_{\mathbb{R}^N}g^2(u)|\nabla u|^2dx = \int_{\mathbb{R}^N}g^2(G^{-1}(v))|\nabla G^{-1}(v)|^2dx = \int_{\mathbb{R}^N}|\nabla v|^2dx < \infty. $ |
After the change of variables, (1.1) changes to
$ −Δv+V(x)G−1(v)g(G−1(v))−K(x)f(G−1(v))g(G−1(v))−λW(x)h(G−1(v))g(G−1(v))=0. $
|
(2.4) |
One can easily derive that if $ v\in X $ is a classical solution to (2.4), then $ u = G^{-1}(v)\in X $ is a classical solution to (1.1). Thus, we only need to seek weak solutions to (2.4). The associated function to (2.4) is
$ Jλ(v)=12∫RN|∇v|2dx+12∫RNV(x)|G−1(v)|2dx−∫RNK(x)F(G−1(v))dx−λ∫RNW(x)H(G−1(v))dx. $
|
(2.5) |
By the conditions on $ g $, $ f $ and $ h $, it is easy to prove that $ J_\lambda $ is well defined and belongs to $ C^1 $ on $ X $. Hence, $ X $ is a proper working space for the problem. Here, we say that $ v\in X $ is a weak solution to (2.4) if
$ ⟨J′λ(v),φ⟩=∫RN[∇v∇φ+V(x)G−1(v)g(G−1(v))φ−K(x)f(G−1(v))g(G−1(v))φ−λW(x)h(G−1(v))g(G−1(v))φ]dx=0 $
|
(2.6) |
for all $ \varphi\in X $.
Before proving the main theorem, we show some technical embedding results for possiblely $ p\leq 2 $, which can be used to deal with sublinear problems comparing with Proposition 2.1.
Lemma 2.2. Assume that $ (W_{0}) $–$ (W_{2}) $ hold. Then, $ X $ is continuously embedded in $ L_W^p(\mathbb{R}^N) $ for all $ p\in(1, 2^*/2) $.
Proof. As mentioned in [2], $ W(x) $ satisfies ($K_1$) and ($K_2$) since it satisfies $ (W_{1}) $ and $ (W_{2}) $. It is clearly $ 2p\in(2, 2^*) $ for $ p\in(1, 2^*/2) $. Therefore, Proposition 2.1 shows that $ X $ is compactly embedded in $ L_W^{2p}(\mathbb{R}^N) $ for every $ p\in(1, 2^*/2) $, and, thus, there exists $ \nu_{W, 2p} > 0 $ such that
$ \int_{\mathbb{R}^N}W(x)|u|^{2p}dx\leq \nu_{W, 2p}^{2p}\|u\|^{2p} $ |
for every $ p\in(1, 2^*/2) $. Moreover, since $ W(x)\in L^1(\mathbb{R}^N) $, by Hölder's inequality and $ (W_{0}) $–$ (W_{2}) $, we deduce for any $ u\in X $
$ ∫RNW(x)|u|pdx=∫RNW(x)12W(x)12|u|pdx≤(∫RNW(x)dx)12(∫RNW(x)|u|2pdx)12≤(‖W(x)‖12p1νW,2p)p‖u‖p $
|
(2.7) |
for all $ p\in(1, 2^*/2) $, implying that $ X $ is continuously embedded in $ L_W^p(\mathbb{R}^N) $.
Lemma 2.3. Assume that $ (W_{0}) $–$ (W_{2}) $ hold. Then, $ X $ is compactly embedded in $ L_W^p(\mathbb{R}^N) $ for all $ p\in(1, 2) $, $ N\geq3 $.
Proof. Lemma 2.2 shows that $ X $ is continuously embedded in $ L_W^p(\mathbb{R}^N) $ for every $ p\in(1, 2) $, and $ N\leq4 $ since $ 2\leq2^*/2 $ in this case. For every $ p\in(1, 2) $, fix $ p_0\in(1, p) $ and $ q_0\in(2, 2^*) $. Then, it follows by Hölder's inequality that
$ ‖u‖pW,p≤‖u‖p0(q0−p)q0−p0W,p0‖u‖q0(p−p0)q0−p0W,q0 for all u∈X, $
|
(2.8) |
which implies by Lemma 2.2 and Proposition 2.1 that $ X $ is compactly embedded in $ L_W^p(\mathbb{R}^N) $ for all $ p\in(1, 2) $ and $ N\leq4 $. Moreover, in the case $ N\geq5 $, for every $ p\in[2^*/2, 2) $, we fix $ p_1\in(1, 2^*/2) $ and $ q_1\in (2, 2^*) $. By a similar inequality, we obtain that $ X $ is compactly embedded in $ L_W^p(\mathbb{R}^N) $ for all $ p\in(1, 2) $, $ N\geq5 $.
In conclusion, $ X $ is compactly embedded in $ L_W^p(\mathbb{R}^N) $ for all $ p\in(1, 2) $.
Now we list the main properties of the function $ G^{-1} $ [14,29].
Lemma 2.4. Suppose that $ g $ satisfies $ (g_{0}) $. Then, the function $ G^{-1}\in C^2(\mathbb{R}, \mathbb{R}) $ satisfies the following properties:
$ (g_1) $ $ G^{-1} $ is increasing and $ G $, $ G^{-1} $ are odd functions;
$ (g_2) $ $ 0 < \frac{d}{dt}\big(G^{-1}(t)\big) = \frac{1}{g(G^{-1}(t))}\leq\frac{1}{g(0)} $ for all $ t\in\mathbb{R} $;
$ (g_3) $ $ |G^{-1}(t)|\leq\frac{|t|}{g(0)} $ for all $ t\in\mathbb{R} $;
$ (g_4) $ $ \lim\limits_{t\rightarrow 0}\frac{G^{-1}(t)}{t} = \frac{1}{g(0)} $;
$ (g_5) $ $ 1\leq \frac{tg(t)}{G(t)}\leq 2 $ and $ 1\leq\frac{G^{-1}(t)g(G^{-1}(t))}{t}\leq 2 $ for all $ t\neq 0 $;
$ (g_6) $ $ \frac{G^{-1}(t)}{\sqrt{t}} $ is non-decreasing in $ (0, +\infty) $ and $ |G^{-1}(t)|\leq (2/g_{\infty})^{1/2}\sqrt{|t|} $ for all $ t\in\mathbb{R} $;
$ (g_7) $ The following inequalities hold
$ |G−1(t)|≥{G−1(1)|t|for all |t|≤1,G−1(1)√|t|for all |t|≥1; $
|
$ (g_8) $ $ \frac{t}{g(t)} $ is increasing and $ \big{|}\frac{t}{g(t)}\big{|}\leq\frac{1}{g_{\infty}} $ for all $ t\in\mathbb{R} $;
$ (g_{9}) $ $ [G^{-1}(s-t)]^2\leq 4([G^{-1}(s)]^2+[G^{-1}(t)]^2) $ for all $ s, t\in\mathbb{R} $;
$ (g_{10}) $ $ \lim\limits_{t\rightarrow +\infty}\frac{G^{-1}(t)}{\sqrt{t}} = (\frac{2}{g_{\infty}})^{1/2} $.
Remark 2.1. Define the function $ \Psi $: $ X\rightarrow \mathbb{R} $ by
$ \Psi(v) = \int_{\mathbb{R}^N}\big(|\nabla v|^2+V(x)[G^{-1}(v)]^2\big)dx. $ |
It is easy to verify that it is a $ C^1 $ function on $ X $ by the conditions on $ g $. Moreover, by $ (g_{3}) $ and $ V(x) > 0\ {for\ all}\ x\in\mathbb{R}^N $, we have
$ \Psi(v)\leq||v||^2 \ \text{for all}\ v\in X, $ |
and as stated in [1], by $ (g_{3}) $, $ (g_{7}) $ and $ (V_{1}) $, there is a constant $ \xi > 0 $ such that
$ \xi||v||^2\leq\Psi(v)+[\Psi(v)]^{2^*/2} \ \text{for all}\ v\in X. $ |
Throughout this paper, $ C $ denotes the various positive constant. $ \nu_{K, q} > 0 $ denotes the Sobolev embedding constant for $ X\hookrightarrow L_K^q(\mathbb{R}^N) $, that is $ \|u\|_{K, q}\leq \nu_{K, q}\|u\| $ for any $ u\in X $, and the definition of Sobolev embedding constant for $ X\hookrightarrow L_W^p(\mathbb{R}^N) $ is similar. Besides, it is well known that the embedding $ D^{1, 2}({\mathbb{R}^{N}})\hookrightarrow L^{2^*}(\mathbb{R}^{N}) $ is continuous, i.e., there exists $ \nu_1 > 0 $ such that $ \|u\|_{2^*}\leq \nu_1\|u\|_{D^{1, 2}({\mathbb{R}^{N}})} $ for any $ u\in D^{1, 2}({\mathbb{R}^{N}}) $.
In this section, we first state a version of the mountain pass theorem due to Ambrosetti and Rabinowitz [4], which is an essential tool in this paper, then we show that the function associated to (2.4) possesses a Cerami sequence at the corresponding mountain pass level. Afterward, the boundedness of the Cerami sequence is established.
We recall the definition of Cerami sequence. Let $ X $ be a real Banach space and $ J_\lambda $: $ X\rightarrow \mathbb{R} $ a functional of class $ C^1 $. We say that $ \{v_n\}\subset X $ is a Cerami sequence at $ c $ ($ (Ce)_c $ for short) for $ J_\lambda $ if $ \{v_n\} $ satisfies
$ Jλ(vn)→c $
|
(3.1) |
and
$ (1+||vn||)J′λ(vn)→0 $
|
(3.2) |
as $ n\rightarrow \infty $. $ J_\lambda $ is said to satisfy the Cerami condition at $ c $, if any Cerami sequence at $ c $ possesses a convergent subsequence.
Theorem 3.1. [31] Let $ X $ be a real Banach space and $ J\in C^1(X, \mathbb{R}) $. Let $ \Sigma $ be a closed subset of $ X $, which disconnects (arcwise) $ X $ into distinct connected $ X_1 $ and $ X_2 $. Suppose further that $ J(0) = 0 $ and
$ (J_1) $ $ 0\in X_1 $, and there is $ \alpha > 0 $ such that $ J|_{\Sigma}\geq\alpha > 0 $,
$ (J_2) $ there is $ e\in X_2 $ such that $ J(e) < 0 $.
Then, $ J $ possesses a $ (Ce)_c $ sequence with $ c\geq\alpha > 0 $ given by
$ c:=infγ∈Λmax0≤t≤1J(γ(t)), $
|
where
$ \Lambda = \{\gamma\in C([0, 1], X):\gamma(0) = 0, \ J(\gamma(1)) < 0\}. $ |
Lemma 3.2. Assume that $ (V, K)\in\mathcal{K} $. $ (g_{0}), $ $ (F_{0}) $–$ (F_{3}), $ $ (W_{0}) $–$ (W_{2}), $ $ (H_{0}) $ and $ (H_{1}) $ hold. Then, there exists $ \lambda_0, \alpha_0 > 0 $ such that for any $ \lambda\in(0, \lambda_0), $ $ J_\lambda $ possesses a Cerami sequence at
$ cλ:=infγ∈Λλmax0≤t≤1Jλ(γ(t))≥α0>0, $
|
where
$ \Lambda_\lambda = \{\gamma\in C([0, 1], X):\gamma(0) = 0, \ J_\lambda(\gamma(1)) < 0\}. $ |
Proof. It is enough to prove that the function satisfies the mountain pass geometry. We only consider the case where $(K_2)$ holds and the proof is similar if $(K_3)$ holds.
First note that $ J_\lambda(0) = 0 $ for any $ \lambda > 0 $. For every $ \rho > 0 $, define
$ {\Sigma}_{\rho}: = \big{\{}v\in X:\int_{\mathbb{R}^N}\big(|\nabla v|^2+V(x)[G^{-1}(v)]^2\big)dx = \rho^2\big{\}}. $ |
Since the function $ \int_{\mathbb{R}^N}\big(|\nabla v|^2+V(x)[G^{-1}(v)]^2\big)dx $ is continuous on $ X $, $ {\Sigma}_\rho $ is a closed subset in $ X $ which disconnects the space $ X $.
$ (1) $ There exists $ \lambda_0, \rho_0, \alpha_0 > 0 $ such that $ J_\lambda(v)\geq\alpha_0 > 0 $ for any $ \lambda\in(0, \lambda_0), \ v\in{\Sigma}_{\rho_0} $. Indeed, for every $ \rho > 0 $, by $(K_2)$, we have
$ ∫RNK(x)|G−1(v)|2dx≤esssupx∈RN|K(x)V(x)|∫RNV(x)|G−1(v)|2dx≤esssupx∈RN|K(x)V(x)|ρ2 $
|
(3.3) |
for any $ v\in {\Sigma}_{\rho} $. Moreover, by $ K(x)\in L^{\infty}(\mathbb{R}^N) $, $ (g_{6}) $ and Sobolev embedding, we conclude that
$ ∫RNK(x)|G−1(v)|22∗dx≤esssupx∈RN|K(x)|∫RN|2g∞v|2∗dx≤esssupx∈RN|K(x)|(ν12g∞)2∗(∫RN|∇v|2dx)2∗2≤esssupx∈RN|K(x)|(ν12g∞)2∗ρ2∗ $
|
(3.4) |
for any $ v\in {\Sigma}_{\rho} $. Thus, by $ (F_{0}) $–$ (F_{2}) $, (3.3) and (3.4), we obtain for any $ \varepsilon > 0 $, there exists $ C_\varepsilon > 0 $ such that
$ ∫RNK(x)F(G−1(v))dx≤ε∫RNK(x)|G−1(v)|2dx+Cε∫RNK(x)|G−1(v)|22∗dx≤εesssupx∈RN|K(x)V(x)|ρ2+Cεesssupx∈RN|K(x)|(ν12g∞)2∗ρ2∗ $
|
(3.5) |
for any $ v\in {\Sigma}_{\rho} $.
In addition, according to Lemma 2.3, $ (g_{2}) $ and $ (g_{3}) $, we deduce that
$ ∫RNW(x)|G−1(v)|τ1dx≤ντ1W,τ1‖G−1(v)‖τ1≤ντ1W,τ1ρτ1 $
|
(3.6) |
and
$ ∫RNW(x)|G−1(v)|τ2dx≤ντ2W,τ2ρτ2 $
|
(3.7) |
for any $ v\in {\Sigma}_{\rho} $.
Thus, by $ (H_{0}) $, $ (H_{1}) $, (3.6) and (3.7), it follows that
$ ∫RNW(x)H(G−1(v))dx≤b1τ1∫RNW(x)|G−1(v)|τ1dx+b2τ2∫RNW(x)|G−1(v)|τ2dx≤b1τ1ντ1W,τ1ρτ1+b2τ2ντ2W,τ2ρτ2 $
|
(3.8) |
for any $ v\in {\Sigma}_{\rho} $.
Choose $ \varepsilon_0 > 0 $ such that $ \text{ess}\sup_{x\in\mathbb{R}^N}\big|\frac{K(x)}{V(x)}\big|\varepsilon_0 < \frac{1}{2} $. By (3.5) and (3.8), we conclude that
$ Jλ(v)≥ρ2(12−esssupx∈RN|K(x)V(x)|ε0−Cε0esssupx∈RN|K(x)|(ν12g∞)2∗ρ2∗−2)−λ(b1τ1ντ1W,τ1ρτ1+b2τ2ντ2W,τ2ρτ2) $
|
for any $ \lambda > 0 $, $ \rho > 0 $, $ v\in {\Sigma}_{\rho} $.
Choose $ \rho_0 > 0 $ such that
$ \frac{1}{2}-\text{ess}\sup\limits_{x\in\mathbb{R}^N}\big|\frac{K(x)}{V(x)}\big|\varepsilon_0-C_{\varepsilon_0}\text{ess}\sup\limits_{x\in\mathbb{R}^N}|K(x)|(\nu_1\frac{2}{g_\infty})^{2^*}\rho_0^{2^*-2} > 0 $ |
and set
$ λ0:=ρ20(12−esssupx∈RN|K(x)V(x)|ε0−Cε0esssupx∈RN|K(x)|(ν12g∞)2∗ρ2∗−20)2(b1τ1ντ1W,τ1ρτ10+b2τ2ντ2W,τ2ρτ20)>0, $
|
$ α0:=ρ202(12−esssupx∈RN|K(x)V(x)|ε0−Cε0esssupx∈RN|K(x)|(ν12g∞)2∗ρ2∗−20)>0. $
|
Then,
$ Jλ(v)≥ρ20(12−esssupx∈RN|K(x)V(x)|ε0−Cε0esssupx∈RN|K(x)|(ν12g∞)2∗ρ2∗−20)−λ(b1τ1ντ1W,τ1ρτ10+b2τ2ντ2W,τ2ρτ20)≥α0>0 $
|
for any $ \lambda\in(0, \lambda_0) $, $ v\in {\Sigma}_{\rho_0} $.
$ (2) $ For any $ \lambda\in(0, \lambda_0) $, there exists $ e\in X $ such that
$ \int_{\mathbb{R}^N}\big(|\nabla e|^2+V(x)|G^{-1}(e)|^2\big)dx > \rho_0 $ |
and $ J_\lambda(e) < 0 $. To this end, for any $ \lambda\in(0, \lambda_0) $, fixed $ v\in X $ is a nonnegative smooth function with $ m(\text{supp}v) > 0 $, where
$ \text{supp}v = \overline{\{x\in\mathbb{R}^N|v(x)\neq0\}} $ |
is the support of $ v $. We prove $ J_\lambda(tv) < 0 $ if $ t > 0 $ and $ \int_{\mathbb{R}^N}\big(|\nabla(tv)|^2+V(x)|G^{-1}(tv)|^2\big)dx $ is large enough. Suppose by contradiction that there exists a sequence $ \{t_n\}\subset\mathbb{R}^+ $ such that
$ \int_{\mathbb{R}^N}\big(|\nabla(t_nv)|^2+V(x)[G^{-1}(t_nv)]^2\big)dx\rightarrow \infty\ \text{as}\ n\rightarrow \infty $ |
and $ J_\lambda(t_nv)\geq0 $ for all $ n\in\mathbb{N} $. By $ (g_{3}) $, we know
$ |t_n|^2\int_{\mathbb{R}^N}\big(|\nabla v|^2+V(x)|v|^2\big)dx\geq\int_{\mathbb{R}^N}\big(|\nabla(t_nv)|^2+V(x)[G^{-1}(t_nv)]^2\big)dx, $ |
which means that $ t_n\rightarrow +\infty $. Set $ \varpi = \frac{v}{\|v\|} $. Noticing that $ K(x), \ W(x) > 0 $, $ \forall x\in\mathbb{R}^N $, by $ (H_{0}) $, $ (F_{0}) $ and $ (g_{3}) $ we get
$ 0≤Jλ(tnv)∫RN(|∇(tnv)|2+V(x)[G−1(tnv)]2)dx≤12−∫suppvK(x)F(G−1(tnv))|G−1(tnv)|4|G−1(tnv)|4|(tnv)|2|ϖ|2dx. $
|
(3.9) |
Since $ t_nv(x)\rightarrow +\infty $ as $ n\rightarrow +\infty $, for $ x\in\text{supp}v $, it follows from $ (g_{10}) $, $ K(x) > 0 $, $ (F_{0}) $, $ (F_{3}) $ and Fatou's lemma that
$ ∫suppvK(x)F(G−1(tnv))|G−1(tnv)|4|G−1(tnv)|4|(tnv)|2|ϖ|2dx→+∞ $
|
as $ n\rightarrow +\infty $, which is a contradiction by inequality (3.9).
The proof is ended.
We now show the boundedness of the Cerami sequence.
Lemma 3.3. Assume that $ (g_{0}), $ $ (V_{1}), $ $ (W_{0}) $–$ (W_{2}), $ $ (H_{0}), $ $ (H_{1}) $ and $ (F_{4}) $ hold, then any $ (Ce)_{c_\lambda} $ sequence of $ J_\lambda $ is bounded in $ X $ for any $ \lambda\in(0, \lambda_0) $.
Proof. Let $ \{v_n\} $ be the corresponding $ (Ce)_{c_\lambda} $ sequence for $ J_\lambda $. Denote $ \omega_n = G^{-1}(v_n)g(G^{-1}(v_n)) $. Then, it follows from (1.7) that
$ ⟨J′λ(vn),ωn⟩≤(1+β)∫RN|∇vn|2dx+∫RNV(x)|G−1(vn)|2dx−∫RNK(x)f(G−1(vn))G−1(vn)dx−λ∫RNW(x)h(G−1(vn))G−1(vn)dx. $
|
(3.10) |
By (1.7) and $ (g_{5}) $, we get
$ |\nabla \omega_n|\leq 2|\nabla v_n|\ \ \ \text{and}\ \ \ |\omega_n|\leq 2|v_n|. $ |
Hence, $ \omega_n\in X $ and $ \|\omega_n\|\leq4\|v_n\| $, which gives
$ |⟨J′λ(vn),ωn⟩|≤J′λ(vn)(1+4‖vn‖)=on(1). $
|
(3.11) |
Therefore, taking into account $ (H_{0}) $, $ (H_{1}) $, $ (W_{0}) $, $ (F_{4}) $, (3.10) and (3.11), we conclude that
$ cλ+on(1)≥Jλ(vn)−1μ⟨J′λ(vn),ωn⟩≥(12−1+βμ)∫RN|∇vn|2dx+(12−1μ)∫RNV(x)[G−1(vn)]2dx−λ∫RNW(x)[b1τ1|G−1(vn)|τ1+b2τ2|G−1(vn)|τ2]dx. $
|
(3.12) |
Hence, combining with $ (W_{0}) $–$ (W_{2}) $, Lemma 2.3, (3.12) and $ (g_{2}) $, we deduce that for any $ \lambda > 0 $,
$ (12−1+βμ)Ψ(vn)≤cλ+λ∫RNW(x)[b1τ1|G−1(vn)|τ1+b2τ2|G−1(vn)|τ2]dx+on(1)≤cλ+λb1ντ1W,τ1τ1‖G−1(vn)‖τ1+λb2ντ2W,τ2τ2‖G−1(vn)‖τ2+on(1)≤cλ+λb1ντ1W,τ1τ1Ψ(vn)τ12+λb2ντ2W,τ2τ2Ψ(vn)τ22+on(1). $
|
Since $ \tau_1, \ \tau_2\in(1, 2) $, $ \{\Psi(v_n)\} $ is bounded in $ X $, by Remark 2.1 we obtain that $ \{v_n\} $ is bounded in $ X $.
Under the hypotheses of Lemmas 3.2 and 3.3, for any fixed $ \lambda\in(0, \lambda_0) $, let $ \{v_n\} $ be the $ (Ce)_{c_\lambda} $ sequence for $ J_\lambda $. Then, by Lemma 3.3 we know that $ \{v_n\} $ is bounded in $ X $. Thus, there exists a subsequence still denoted by $ \{v_n\} $, and $ v\in X $ such that
$ vn⇀v in X, vn→v in Lsloc(RN) for any s∈[1,2∗) and vn→v a.e., on RN, $
|
(4.1) |
and there is $ L > 0 $ such that
$ ∫RN|∇vn|2dx+∫RNV(x)|vn|2dx≤L and ∫RN|vn|2∗dx≤L, ∀n∈N. $
|
(4.2) |
We conclude this section showing that the weak limit $ v $ is a positive solution to (1.1).
Lemma 4.1. Assume that $ (g_{0}), $ $ (W_{0}) $–$ (W_{2}), $ $ (H_{0}), $ $ (H_{1}) $ hold and $ \{v_n\} $ is a $ (Ce)_{c_\lambda} $ sequence for $ J_\lambda $ given by Lemmas 3.2 and 3.3. Then, the following statements hold:
$ limn→+∞∫RNW(x)H(G−1(vn))dx=∫RNW(x)H(G−1(v))dx, $
|
(4.3) |
$ limn→+∞∫RNW(x)h(G−1(vn))g(G−1(vn))φdx=∫RNW(x)h(G−1(v))g(G−1(v))φdx, for any φ∈X, $
|
(4.4) |
$ limn→+∞∫RNW(x)h(G−1(vn))G−1(vn)dx=∫RNW(x)h(G−1(v))G−1(v)dx, $
|
(4.5) |
$ limn→+∞∫RNW(x)h(G−1(vn))g(G−1(vn))vndx=∫RNW(x)h(G−1(v))g(G−1(v))vdx. $
|
(4.6) |
Proof. First, we give the proof of (4.3). Since $ \tau_1, \ \tau_2\in(1, 2) $, from $ (W_{0}) $–$ (W_{2}) $ and Lemma 2.3, we have
$ ∫RNW(x)|vn|τ1dx→∫RNW(x)|v|τ1dx and ∫RNW(x)|vn|τ2dx→∫RNW(x)|v|τ2dx. $
|
(4.7) |
Then, given $ \varepsilon > 0 $, there is $ r > 0 $ such that
$ ∫BcrW(x)|vn|τ1dx<ε and ∫BcrW(x)|vn|τ2dx<ε for all n∈N, $
|
(4.8) |
where $ B_r^c: = \{x\in\mathbb{R}^N:|x| > r\} $, which together with $ (H_{0}) $, $ (H_{1}) $ and $ (g_{3}) $ yields that
$ ∫BcrW(x)H(G−1(vn))dx≤b1τ1∫BcrW(x)|G−1(vn)|τ1dx+b2τ2∫BcrW(x)|G−1(vn)|τ2dx≤b1τ1∫BcrW(x)|vn|τ1dx+b2τ2∫BcrW(x)|vn|τ2dx<(b1τ1+b2τ2)ε $
|
for any $ n\in\mathbb{N} $.
Moreover, for each fixed $ r > 0 $, it is easy to verify that
$ limn→+∞∫Br(0)W(x)H(G−1(vn))dx=∫Br(0)W(x)H(G−1(v))dx, $
|
where $ B_r(0) = \{x\in\mathbb{R}^N:|x|\leq r\} $. This completes the proof of (4.3).
Proof. Now we are going to prove (4.4). Noticing (4.7), given $ \varepsilon > 0 $, there is $ r > 0 $ such that
$ ∫BcrW(x)|vn|τ1dx<ετ1τ1−1 and ∫BcrW(x)|vn|τ2dx<ετ2τ2−1 for all n∈N. $
|
(4.9) |
By $ (H_{0}) $, $ (H_{1}) $, $ (W_{0}) $, $ (g_{2}) $, $ (g_{3}) $ and Hölder's inequality, we obtain that
$ |∫BcrW(x)h(G−1(vn))g(G−1(vn))φdx|≤b1∫BcrW(x)|G−1(vn)|τ1−1g(G−1(vn))|φ|dx+b2∫BcrW(x)|G−1(vn)|τ2−1g(G−1(vn))|φ|dx≤b1∫BcrW(x)|vn|τ1−1|φ|dx+b2∫BcrW(x)|vn|τ2−1|φ|dx≤b1(∫BcrW(x)|vn|τ1dx)τ1−1τ1(∫BcrW(x)|φ|τ1dx)1τ1+b2(∫BcrW(x)|vn|τ2dx)τ2−1τ2(∫BcrW(x)|φ|τ2dx)1τ2 $
|
(4.10) |
for any $ n\in\mathbb{N} $, $ \varphi\in X $. Since $ \tau_1, \tau_2\in(1, 2) $, Lemma 2.3 implies that $ \int_{B_r^c}W(x)|\varphi|^{\tau_1}dx < \infty $ and $ \int_{B_r^c}W(x)|\varphi|^{\tau_2}dx < \infty $. Thus, combining with (4.9) and (4.10), we conclude that
$ |∫BcrW(x)h(G−1(vn))g(G−1(vn))φdx|<C1ε $
|
for any $ \varphi\in X $, where $ C_1 = b_1\|\varphi\|_{W, \tau_1}+b_2\|\varphi\|_{W, \tau_2} $.
Moreover, for each fixed $ r > 0 $, it is easy to verify that
$ limn→+∞∫Br(0)W(x)h(G−1(vn))g(G−1(vn))φdx=∫Br(0)W(x)h(G−1(v))g(G−1(v))φdx for any φ∈X. $
|
This completes the proof of (4.4).
Repeating the similar arguments used in the proofs of (4.3) and (4.4), we can obtain that (4.5) and (4.6) hold.
Lemma 4.2. Assume that $ (V, K)\in\mathcal{K}, $ $ (g_{0}), $ $ (F_{0}) $–$ (F_{2}) $ hold and $ \{v_n\} $ is a $ (Ce)_{c_\lambda} $ sequence for $ J_\lambda $ given by Lemmas 3.2 and 3.3. Then, the following statements hold:
$ limn→+∞∫RNK(x)F(G−1(vn))dx=∫RNK(x)F(G−1(v))dx, $
|
(4.11) |
$ limn→+∞∫RNK(x)f(G−1(vn))g(G−1(vn))φdx=∫RNK(x)f(G−1(v))g(G−1(v))φdx for all φ∈X, $
|
(4.12) |
$ limn→+∞∫RNK(x)f(G−1(vn))G−1(vn)dx=∫RNK(x)f(G−1(v))G−1(v)dx, $
|
(4.13) |
$ limn→+∞∫RNK(x)f(G−1(vn))g(G−1(vn))vndx=∫RNK(x)f(G−1(v))g(G−1(v))vdx. $
|
(4.14) |
Proof. (1) We begin the proof of (4.11) by assuming that $(K_2)$ holds. By $ (F_{0}) $–$ (F_{2}) $, we obtain that there exists $ C_2 > 0 $ such that
$ F(G−1(s))≤C2|G−1(s)|2+C2|G−1(s)|22∗ for all s∈R, $
|
which together with $ (F_{0}) $–$ (F_{2}) $, $ (g_{3}) $ and $ (g_{6}) $ yields that, for any fixed $ q\in(2, 2^*) $, given $ \varepsilon > 0 $ there exists $ 0 < s_0 < s_1 $ such that
$ |F(G−1(s))|≤ε2|G−1(s)|2+ε22∗|G−1(s)|22∗+χ[s0,s1](|θ|)(C2|G−1(s)|2+C2|G−1(s)|22∗)≤ε2|G−1(s)|2+ε22∗|G−1(s)|22∗+χ[s0,s1](|θ|)C2(1|s0|q−2+|s1|22∗−q)|G−1(s)|q≤ε(12+(2/g∞)2∗22∗)(|s|2+|s|2∗)+C2(1|s0|q−2+|s1|22∗−q)|s|q $
|
(4.15) |
for all $ s\in\mathbb{R} $, where $ \theta = G^{-1}(s) $.
In addition, by $ K(x), \ V(x) > 0 $ for all $ x\in\mathbb{R}^N $, $(K_2)$ and $ K(x)\in L^{\infty}(\mathbb{R}^N) $, we obtain that there exists $ C_3 > 0 $ such that
$ ∫RN(K(x)|s|2+K(x)|s|2∗)dx≤∫RN(esssupx∈RN|K(x)V(x)|V(x)|s|2+esssupx∈RN|K(x)||s|2∗)dx≤C3∫RN(V(x)|s|2+|s|2∗)dx $
|
(4.16) |
for any $ s\in\mathbb{R}^N $.
Furthermore, noticing $ q\in(2, 2^*) $, then from Proposition 2.1 we have
$ ∫RNK(x)|vn|qdx→∫RNK(x)|v|qdx as n→+∞, $
|
(4.17) |
which gives that there is $ r > 0 $ such that
$ ∫BcrK(x)|vn|qdx<εC2(1|s0|q−2+|s1|22∗−q), ∀n∈N. $
|
(4.18) |
Therefore, combining with (4.2), (4.15), (4.16) and (4.18), we conclude that
$ |∫BcrK(x)F(G−1(vn))dx|≤ε(12+(2/g∞)2∗22∗)∫RN(K(x)|vn|2+K(x)|vn|2∗)dx+C2(1|s0|q−2+|s1|22∗−q)∫BcrK(x)|vn|qdx≤ε(12+(2/g∞)2∗22∗)C3∫RN(V(x)|vn|2+|vn|2∗)dx+C2(1|s0|q−2+|s1|22∗−q)∫BcrK(x)|vn|qdx<(C3L2+(2/g∞)2∗22∗C3L+1)ε $
|
for all $ n\in\mathbb{N} $.
Now, if $(K_3)$ holds, by $ (F_{0}) $–$ (F_{2}) $, we obtain there exists $ C_4 > 0 $ such that
$ F(G−1(s))≤C4|G−1(s)|σ+C4|G−1(s)|22∗, $
|
which together with $ (F_{0}) $–$ (F_{2}) $, $ (g_{3}) $ and $ (g_{6}) $ yields that, given $ \varepsilon\in(0, 1) $, there exists $ 0 < s_0 < s_1 $ such that
$ F(G−1(s))≤εσ|G−1(s)|σ+ε22∗|G−1(s)|22∗+χ[s0,s1](|θ|)(C4|G−1(s)|σ+C4|G−1(s)|22∗)≤εσ|G−1(s)|σ+ε22∗|G−1(s)|22∗+χ[s0,s1](|θ|)C4(1+|s1|22∗−σ)|G−1(s)|σ≤1σ|s|σ+ε(2/g∞)2∗22∗|s|2∗+C4(1+|s1|22∗−σ)|s|σ≤(1σ+C4+C4|s1|22∗−σ)|s|σ+ε(2/g∞)2∗22∗|s|2∗ $
|
(4.19) |
for any $ s\in\mathbb{R} $.
Furthermore, noticing $ \sigma\in(2, 2^*) $, by Proposition 2.1 we have
$ ∫RNK(x)|vn|σdx→∫RNK(x)|v|σdx as n→+∞, $
|
(4.20) |
which gives that there is $ r > 0 $ such that
$ ∫BcrK(x)|vn|σdx<ε1σ+C4+C4|s1|22∗−σ, ∀n∈N. $
|
(4.21) |
Therefore, by (4.2), (4.19), (4.21), $ K(x) > 0 $ for all $ x\in\mathbb{R}^N $ and $ K\in L^\infty(\mathbb{R}^N) $, we obtain that
$ ∫BcrK(x)F(G−1(vn))dx≤(1σ+C4+C4|s1|22∗−σ)∫BcrK(x)|vn|σdx+ε(2/g∞)2∗22∗∫BcrK(x)|vn|2∗dx≤(1σ+C4+C4|s1|22∗−σ)∫BcrK(x)|vn|σdx+ε(2/g∞)2∗22∗esssupx∈RN|K(x)|∫Bcr|vn|2∗dx<C5ε $
|
for any $ n\in\mathbb{N} $, where
$ C_5 = \big(1+\frac{(2/g_\infty)^{2^*}}{22^*}\text{ess}\sup\limits_{x\in\mathbb{R}^N}|K(x)|L\big). $ |
Furthermore, for each fixed $ r > 0 $, it is easy to verify that
$ limn→+∞∫Br(0)K(x)F(G−1(vn))dx=∫Br(0)K(x)F(G−1(v))dx. $
|
This completes the proof of (4.11).
(2) We begin the proof of (4.12) if $(K_2)$ holds. By $ (F_{0}) $–$ (F_{2}) $, we have that there is $ C_6 > 0 $ such that
$ f(G−1(s))g(G−1(s))≤C6|G−1(s)|g(G−1(s))+C6|G−1(s)|22∗−1g(G−1(s)) for all s∈R, $
|
which together with $ (F_{0}) $–$ (F_{2}) $, $ (g_{2}) $, $ (g_{3}) $, $ (g_{6}) $ and $ (g_{8}) $ yields that, for any fixed $ q\in(2, 2^*) $, given $ \varepsilon\in(0, 1) $ there exists $ 0 < s_0 < s_1 $ such that
$ f(G−1(s))g(G−1(s))≤ε|G−1(s)|g(G−1(s))+ε|G−1(s)|22∗−1g(G−1(s))+χ[s0,s1](|θ|)C6(1|s0|q−2+|s1|22∗−q)|G−1(s)|q−1≤ε|G−1(s)|+ε1g∞|G−1(s)|22∗−2+C6(1|s0|q−2+|s1|22∗−q)|G−1(s)|q−1≤ε(1+(2/g∞)2∗)(|s|+|s|2∗−1)+C6(1|s0|q−2+|s1|22∗−q)|s|q−1 $
|
(4.22) |
for any $ s\in\mathbb{R} $.
On the other hand, noticing $ q\in(2, 2^*) $, by (4.17) we have there is $ r > 0 $ such that
$ ∫BcrK(x)|vn|qdx<εC6(1|s0|q−2+|s1|22∗−q), ∀n∈N. $
|
(4.23) |
Thus, taking into account $(K_2)$, $ K(x), V(x) > 0 $ for all $ x\in\mathbb{R}^N $, $ K(x)\in L^{\infty}(\mathbb{R}^N) $, (4.22) and Hölder's inequality, we obtain that, for any $ \varphi\in X $,
$|∫BcrK(x)f(G−1(vn))g(G−1(vn))φdx|≤ε(1+(2/g∞)2∗)(∫BcrK(x)|vn||φ|dx+∫BcrK(x)|vn|2∗−1|φ|dx)+C6(1|s0|q−2+|s1|22∗−q)∫BcrK(x)|vn|q−1|φ|dx≤ε(1+(2/g∞)2∗)[(∫Bcresssupx∈RN|K(x)V(x)|V(x)|vn|2dx)12(∫BcrK(x)|φ|2dx)12+esssupx∈RN|K(x)|(∫Bcr|vn|2∗dx)2∗−12∗(∫Bcr|φ|2∗dx)12∗]+C6(1|s0|q−2+|s1|22∗−q)(∫BcrK(x)|vn|qdx)q−1q(∫BcrK(x)|φ|qdx)1q $
|
(4.24) |
for any $ n\in\mathbb{N} $. Moreover, for any $ \varphi\in X $, by Proposition 2.1, $(K_2)$ and Sobolev embedding, we have
$ \int_{\mathbb{R}^N}K(x)|\varphi|^qdx < +\infty, \ \ \int_{\mathbb{R}^N}K(x)|\varphi|^2dx < \text{ess}\sup\limits_{x\in\mathbb{R}^N}\big|\frac{K(x)}{V(x)}\big|\int_{\mathbb{R}^N}V(x)|\varphi|^2dx < +\infty $ |
and
$ \int_{\mathbb{R}^N}|\varphi|^{2^*}dx < \nu_1\|\varphi\|^{2^*} < +\infty, $ |
respectively. Thus, it follows from (4.2), (4.23) and (4.24) that
$ |∫BcrK(x)f(G−1(vn))g(G−1(vn))φdx|<C7ε, $
|
where
$ C_7 = \big(1+(2/g_\infty)^{2^*}\big)\big[\big(\text{ess}\sup\limits_{x\in\mathbb{R}^N}\big|\frac{K(x)}{V(x)}\big|\big)L^{\frac{1}{2}}\|\varphi\| +\text{ess}\sup\limits_{x\in\mathbb{R}^N}|K(x)|L^{\frac{2^*-1}{2^*}}\|\varphi\|_{2^*}\big]+\|\varphi\|_{K, q}. $ |
Now, if $(K_3)$ holds, by $ (F_{0}) $–$ (F_{2}) $ we obtain there exists $ C_8 > 0 $ such that
$ f(G−1(s))g(G−1(s))≤C8|G−1(s)|σ−1g(G−1(s))+C8|G−1(s)|22∗−1g(G−1(s)), $
|
which together with $ (F_{0}) $–$ (F_{2}) $, $ (g_{3}) $, $ (g_{6}) $ and $ (g_{8}) $ yields that, given $ \varepsilon\in(0, 1) $ there exists $ 0 < s_0 < s_1 $ such that
$ f(G−1(s))g(G−1(s))≤ε|G−1(s)|σ−1g(G−1(s))+ε|G−1(s)|22∗−1g(G−1(s))+χ[s0,s1](|θ|)(C8|G−1(s)|σ−1g(G−1(s))+C8|G−1(s)|22∗−1g(G−1(s)))≤|G−1(s)|σ−1+ε(1/g∞)|G−1(s)|22∗−2+χ[s0,s1](|θ|)C8(1+|s1|22∗−σ)|G−1(s)|σ−1≤(1+C8+C8|s1|22∗−σ)|s|σ−1+ε(2/g∞)2∗|s|2∗−1 $
|
(4.25) |
for any $ s\in\mathbb{R} $. Furthermore, from $ (V, K)\in\mathcal{K} $ and Proposition 2.1, we infer that there is $ r > 0 $ such that
$ ∫BcrK(x)|vn|σdx<εσσ−1(1+C8+C8|s1|22∗−σ)σσ−1, ∀n∈N. $
|
(4.26) |
Combining with Hölder's inequality, (4.2), (4.25), (4.26), $ K(x), V(x) > 0 $ for all $ x\in\mathbb{R}^N $, $ K(x)\in L^{\infty}(\mathbb{R}^N) $ and Proposition 2.1, it follows that
$ |∫BcrK(x)f(G−1(vn))g(G−1(vn))φdx|≤(1+C8+C8|s1|22∗−σ)∫BcrK(x)|vn|σ−1|φ|dx+ε(2/g∞)2∗∫BcrK(x)|vn|2∗−1|φ|dx≤(1+C8+C8|s1|22∗−σ)(∫BcrK(x)|vn|σdx)σ−1σ(∫BcrK(x)|φ|σdx)1σ+ε(2/g∞)2∗esssupx∈RN|K(x)|(∫Bcr|vn|2∗dx)2∗−12∗(∫Bcr|φ|2∗dx)12∗<C9ε, $
|
where
$ C_{9} = \|\varphi\|_{K, \sigma}+(2/g_\infty)^{2^*}L^{\frac{2^*-1}{2^*}}\|\varphi\|_{2^*}\text{ess}\sup\limits_{x\in\mathbb{R}^N}|K(x)| $ |
for all $ \varphi\in X $, $ n\in\mathbb{N} $.
Moreover, for each fixed $ r > 0 $, it is easy to verify that
$ limn→+∞∫Br(0)K(x)f(G−1(vn))g(G−1(vn))φdx=∫Br(0)K(x)f(G−1(v))g(G−1(v))φdx. $
|
This completes the proof of (4.12)
Repeating the similar arguments used in the proof of (4.11) and (4.12), we obtain that (4.13) and (4.14) hold.
Lemma 4.3. Assume that $ (V, K)\in\mathcal{K}, $ $ (g_{0}), $ $ (V_{1}), $ $ (W_{0}) $–$ (W_{2}), $ $ (F_{0}) $–$ (F_{2}), $ $ (H_{0}), $ $ (H_{1}) $ hold and $ \{v_n\} $ is a $ (Ce)_{c_\lambda} $ sequence for $ J_\lambda $ given by Lemmas 3.2 and 3.3. Then, the following statements hold:\\ (i) For each $ \varepsilon > 0 $ there exists $ r_0 > 1 $, such that for any $ r > r_0 $
$ lim supn→+∞∫Bc2r(|∇vn|2+V(x)|G−1(vn)|2)dx<(3+λ)ε, $
|
(4.27) |
$ lim supn→+∞∫Bc2r(|∇vn|2+V(x)G−1(vn)g(G−1(vn))vn)dx<(3+λ)ε $
|
(4.28) |
and
$ limn→+∞∫RNV(x)|G−1(vn)|2dx=∫RNV(x)|G−1(v)|2dx, $
|
(4.29) |
$ limn→+∞∫RNV(x)G−1(vn)g(G−1(vn))vndx=∫RNV(x)G−1(v)g(G−1(v))vdx. $
|
(4.30) |
(ii) The weak limit $ v $ of $ \{v_n\} $ is a critical point for the function $ J_\lambda $ on $ X $.
(iii) The weak limit $ v $ is a nontrivial critical point of $ J_\lambda $ and $ J_\lambda(v) = c_\lambda $. Moreover, the function $ J_\lambda $ satisfies the Cerami condition on $ X $.
Proof. (i) For $ r > 1 $, we choose a cut-off function $ \eta = \eta_r\in C_0^\infty(B_r^c) $ such that
$ η≡1 in Bc2r, η≡0 in Br, 0≤η≤1 $
|
(4.31) |
and
$ |∇η|≤2r for all x∈RN. $
|
(4.32) |
As $ \{v_n\} $ is bounded in $ X $, the sequence $ \{\eta\omega_n\} $ where $ \omega_n = G^{-1}(v_n)g(G^{-1}(v_n)) $ is also bounded in $ X $. Hence, from (3.11) we have
$ |\langle J_\lambda'(v_n), \eta\omega_n\rangle| = o_n(1), $ |
that is
$ ∫RN(1+g′(t)|t=G−1(vn)G−1(vn)g(G−1(vn)))η|∇vn|2dx+∫RNV(x)[G−1(vn)]2ηdx=−∫RN∇η∇vnωndx+∫RNK(x)f(G−1(vn))G−1(vn)ηdx +∫RNλW(x)h(G−1(vn))G−1(vn)ηdx+on(1). $
|
(4.33) |
Then, by $ (g_{0}) $, $ (F_{0}) $, $ (H_{0}) $, (4.31) and (4.33) we infer that
$ ∫Bcr(|∇vn|2+V(x)|G−1(vn)|2)ηdx≤on(1)+∫Bcr|∇η||∇vn||ωn|dx+∫BcrK(x)f(G−1(vn))G−1(vn)ηdx+∫BcrλW(x)h(G−1(vn))G−1(vn)ηdx $
|
(4.34) |
for any $ r > 1 $.
By (4.2), (4.32), $ (g_{5}) $ and Hölder's inequality, we obtain
$ ∫Bcr|∇η||∇vn||ωn|dx≤4r∫{r≤|x|≤2r}|∇vn||vn|dx≤4r(∫RN|∇vn|2dx)12(∫{r≤|x|≤2r}|vn|2dx)12≤4rL12(∫{r≤|x|≤2r}|vn|2dx)12 $
|
(4.35) |
for any $ r > 1 $, $ n\in\mathbb{N} $. Noticing that $ v_n\rightarrow v $ in $ L^2(B_{2r}\setminus B_r) $ and $ |B_{2r}\setminus B_r|\leq|B_{2r}| = \omega_N(2r)^N $ for any fixed $ r > 1 $, then (4.35) follows that
$ lim supn→+∞∫Bcr|∇η||∇vn||ωn|dx≤4rL12(∫{r≤|x|≤2r}|v|2dx)12≤4rL12(∫{r≤|x|≤2r}|v|2∗dx)12∗|B2r∖Br|1N,≤8L12ω1NN(∫{r≤|x|≤2r}|v|2∗dx)12∗ $
|
(4.36) |
for any $ r > 1 $. Furthermore, for any $ \varepsilon > 0 $ there exists $ r_1 > 1 $, and for any $ r > r_1 $
$ 8L12ω1NN(∫{r≤|x|≤2r}v2∗dx)12∗<ε. $
|
(4.37) |
Therefore, combining with (4.36) and (4.37), we have that for any $ r > r_1 $,
$ lim supn→+∞∫Bcr|∇η||∇vn||ωn|dx<ε. $
|
(4.38) |
In addition, according to the (4.5) and (4.13), we infer that there is $ r_2 > 1 $ such that
$ ∫BcrλW(x)h(G−1(vn))G−1(vn)ηdx≤∫BcrλW(x)h(G−1(vn))G−1(vn)dx<λε, for any n∈N $
|
(4.39) |
and
$ ∫BcrK(x)f(G−1(vn))G−1(vn)ηdx≤∫BcrK(x)f(G−1(vn))G−1(vn)dx<ε, for any n∈N $
|
(4.40) |
for any $ r > r_2 $.
Set $ r_0 = \max\{r_1, r_2\} $. Then, taking into account (4.34), (4.38), (4.39) and (4.40), we know that (4.27) is valid.
Noticing $ (g_{5}) $, we know that $ \big|\frac{v_n}{g(G^{-1}(v_n))}\big| < \big|G^{-1}(v_n)\big| $ for any $ n\in\mathbb{N} $. Then, (4.27) implies that (4.28) holds.
Moreover, the limit (4.27) gives that
$ lim supn→∞∫Bc2rV(x)|G−1(vn)|2dx<(3+λ)ε $
|
(4.41) |
for any $ r > r_0 $ and consequently,
$ ∫Bc2rV(x)|G−1(v)|2dx<(3+λ)ε $
|
(4.42) |
for any $ r > r_0 $. Since $ v_n\rightarrow v $ in $ L^2(B_{2r}(0)) $ for any fixed $ r\in(0, +\infty) $, then by $ (g_{3}) $ and the continuity of $ V(x) $, using the Lebesgue dominated convergence theorem we know that
$ limn→+∞∫B2r(0)V(x)|G−1(vn)|2dx=∫B2r(0)V(x)|G−1(v)|2dx. $
|
(4.43) |
Then, (4.41)–(4.43) yield that
$ lim supn→+∞|∫RNV(x)[|G−1(vn)|2−|G−1(v)|2]dx|<2(3+λ)ε $
|
and, hence, (4.29) holds. Similarly, it follows from (4.28) that (4.30) holds.
(ii) It is clear that
$ √V(x)G−1(vn(x))g(G−1(vn(x)))→√V(x)G−1(v(x))g(G−1(v(x))) a.e., x∈RN $
|
as $ n\rightarrow +\infty $. Noting that $ \big\{\sqrt{V(x)}\frac{G^{-1}(v_n(x))}{g(G^{-1}(v_n(x)))}\big\} $ is bounded in $ L^2(\mathbb{R}^N) $ and $ \sqrt{V}\varphi\in L^2(\mathbb{R}^N) $ for any $ \varphi\in X $, we have that
$ √V(x)G−1(vn(x))g(G−1(vn(x)))⇀√V(x)G−1(v(x))g(G−1(v(x))) $
|
in $ L^2(\mathbb{R}^N) $, as $ n\rightarrow +\infty $ and, hence, the following equality holds
$ limn→+∞∫RNV(x)[G−1(vn(x))g(G−1(vn(x)))−G−1(v(x))g(G−1(v(x)))]φdx=0, for any φ∈X. $
|
(4.44) |
Furthermore, since $ v_n\rightharpoonup v $ in $ D^{1, 2}(\mathbb{R}^N) $ and $ \varphi\in D^{1, 2}(\mathbb{R}^N) $, we have
$ ∫RN∇vn∇φdx→∫RN∇v∇φdx, for any φ∈X. $
|
(4.45) |
Thus, by (4.4), (4.12), (4.44) and (4.45), we deduce that
$ limn→+∞⟨J′λ(vn),φ⟩=⟨J′λ(v),φ⟩, ∀ φ∈X. $
|
Thus, $ J_\lambda'(v) = 0 $, which implies that (ii) holds.
(iii) We have proved that $ J_\lambda'(v) = 0 $. Now, we show that $ v\neq 0 $. Suppose that $ v\equiv 0 $; because $ \{v_n\} $ is a $ (Ce)_{c_\lambda} $ sequence, according to (3.2), we know
$ (J′λ(vn),vn)=∫RN|∇vn|2dx+∫RNV(x)G−1(vn)g(G−1(vn))vndx−∫RNK(x)f(G−1(vn))g(G−1(vn))vndx−λ∫RNW(x)h(G−1(vn))g(G−1(vn))vndx→0. $
|
(4.46) |
Moreover, by (4.5), (4.13) and (4.30), we get
$ limn→+∞∫RNV(x)G−1(vn)g(G−1(vn))vndx=0, $
|
$ limn→+∞∫RNW(x)h(G−1(vn))G−1(vn)dx=0 $
|
and
$ limn→+∞∫RNK(x)f(G−1(vn))G−1(vn)dx=0. $
|
Then, it follows from (4.46) that
$ ∫RN|∇vn|2dx→0 as n→+∞. $
|
In addition, from (4.3), (4.11) and (4.29), we have
$ limn→+∞∫RNV(x)|G−1(vn)|2dx=0, $
|
$ limn→+∞∫RNW(x)H(G−1(vn))dx=0 $
|
and
$ limn→+∞∫RNK(x)F(G−1(vn))dx=0. $
|
Hence,
$ Jλ(vn)=12∫RN|∇vn|2dx+12∫RNV(x)|G−1(vn)|2dx−∫RNK(x)F(G−1(vn))dx−λ∫RNW(x)H(G−1(vn))dx→0, $
|
which is a contradiction to $ J_\lambda(v_n)\rightarrow c_\lambda > \alpha_0 > 0 $. Therefore, $ v\neq 0 $.
Now, we show that $ J_\lambda(v) = c_\lambda $. By $ \langle J_\lambda'(v_n), v_n\rangle = o_n(1) $, passing to the limit in the following expression
$ ∫RN|∇vn|2dx=−∫RNV(x)G−1(vn)g(G−1(vn))vndx+∫RNK(x)f(G−1(vn))g(G−1(vn))vndx+λ∫RNW(x)h(G−1(vn))g(G−1(vn))vndx+on(1) $
|
and using (4.6), (4.14) and (4.30) together with $ (J_\lambda'(v), v) = 0 $, we obtain that
$ limn→+∞∫RN|∇vn|2dx=∫RN|∇v|2dx. $
|
(4.47) |
By (4.3), (4.11), (4.29) and (4.47), we conclude that
$ Jλ(vn)=12∫RN|∇vn|2dx+12∫RNV(x)|G−1(vn)|2dx −∫RNK(x)F(G−1(vn))dx−λ∫RNW(x)H(G−1(vn))dx→Jλ(v), $
|
which results that $ J_\lambda(v) = c_\lambda $.
To show that the function $ J_\lambda $ satisfies the Cerami condition, we verify that $ \|v_n-v\|\rightarrow 0 $. By Remark 2.1, we have
$ ξ‖vn−v‖2≤Ψ(vn−v)+[Ψ(vn−v)]2∗/2, $
|
where
$ \Psi(v_n-v) = \int_{\mathbb{R}^N}\big[|\nabla(v_n-v)|^2+V(x)|G^{-1}(v_n-v)|^2\big]dx. $ |
Combining with $ (g_{9}) $, (4.41) and (4.42), we obtain that given $ \varepsilon > 0 $
$ lim supn→+∞∫Bc2rV(x)[G−1(vn−v)]2dx≤lim supn→+∞∫Bc2r4V(x)[|G−1(vn)|2+|G−1(v)|2]dx<8(3+λ)ε $
|
(4.48) |
for any $ r > r_0 $. Furthermore, noticing $ v_n\rightarrow v $ in $ L^2(B_{2r}) $ for any fixed $ r > 0 $, by $ (g_{3}) $ we infer that
$ 0≤limn→+∞∫B2rV(x)[G−1(vn−v)]2dx≤limn→+∞∫B2rV(x)|vn−v|2dx=0, $
|
implying that
$ limn→+∞∫B2rV(x)[G−1(vn−v)]2dx=0. $
|
(4.49) |
Thus, (4.48) and (4.49) lead to
$ limn→+∞∫RNV(x)[G−1(vn−v)]2dx=0, $
|
which together with (4.47) yields that $ \Psi(v_n-v)\rightarrow0 $. Then, $ \|v_n-v\|\rightarrow 0 $ holds, implying that $ J_\lambda $ satisfies the Cerami condition.
Proof of Theorem 1.1. Combining all the results above, we get that for every $ \lambda\in(0, \lambda_0) $, (2.4) possesses a nontrivial solution $ v $. Furthermore, letting $ v^{-} = \max\{-v, 0\} $, by $ J_\lambda'(v) = 0 $, $ (F_{0}) $ and $ (W_{0}) $, we have that
$ ⟨J′λ(v),v−⟩=∫RN[|∇v−|2+V(x)G−1(v)g(G−1(v))v−]dx=0. $
|
Since $ G^{-1}(v)v^{-}\geq 0 $, $ V(x) > 0 $ and $ g(G^{-1}(v)) > 0 $, we have that
$ ∫RN|∇v−|2dx=0 and ∫RNV(x)G−1(v)g(G−1(v))v−dx=0. $
|
Thus, $ v^{-} = 0 $ a.e., $ x\in\mathbb{R}^N $. Therefore, $ v $ is a positive solution to (2.4); that is, $ u = G^{-1}(v) $ is a positive solution to (1.1).
Now, we prove Proposition 1.2. At first we show the following lemma.
Set $ S_{r} = \big\{v\in X:\|v\| = r\big\} $ and $ \mathcal{B}_{r} = \big\{v\in X:\|v\| < r\big\} $.
Lemma 4.4. Assume that $ (V, K)\in\mathcal{K} $ where $(K_2)$ holds and $ (g_{0}), $ $ (V_{1}), $ $ (W_{0}) $–$ (W_{2}), $ $ (F'_{0}), $ $ (F_{1}), $ $ (F_{2}), $ $ (F_{4}), $ $ (H'_{0}) $–$ (H'_{2}) $ hold. Then, there exists $ r_1, \alpha_1, \lambda_1 > 0 $ such that $ J_{\lambda}|_{S_{r_1}}\geq\alpha_1 $ and $ \inf\limits_{v\in \mathcal{B}_{r_1}}J_{\lambda}(v) < 0 $ for any $ \lambda\in(0, \lambda_1) $.
Proof. Noticing Remark 2.1, we have
$ Ψ(v)≥ξ||v||2−[Ψ(v)]2∗/2≥ξ||v||2−||v||2∗ for all v∈X. $
|
(4.50) |
Choose $ \varepsilon_1 > 0 $ such that
$ \frac{1}{2}-\frac{\varepsilon_1}{2}\text{ess}\sup\limits_{x\in\mathbb{R}^N}\big|\frac{K(x)}{V(x)}\big| > 0. $ |
By $ (F'_{0}) $, $ (F_{1}) $, $ (F_{2}) $, $ (H'_{0}) $, $ (H'_{1}) $, $ (W_{0}) $–$ (W_{2}) $, (K_2), $ K(x)\in L^{\infty}(\mathbb{R}^N) $, $ K(x), V(x) > 0 $ for all $ x\in\mathbb{R}^N $, $ (g_{6}) $, (4.50) and Lemma 2.3, we obtain there exists $ C_{\varepsilon_1} > 0 $ such that
$Jλ(v)≥12∫RN|∇v|2dx+12∫RNV(x)|G−1(v)|2dx−ε12∫RNK(x)|G−1(v)|2dx−Cε122∗∫RNK(x)|G−1(v)|22∗dx−λb3τ3∫RNW(x)|G−1(v)|τ3dx≥(12−ε12esssupx∈RN|K(x)V(x)|)Ψ(v)−esssupx∈RN|K(x)|(2g∞)2∗Cε122∗∫RN|v|2∗dx−λb3τ3∫RNW(x)|G−1(v)|τ3dx≥(12−ε12esssupx∈RN|K(x)V(x)|)(ξ||v||2−||v||2∗)−esssupx∈RN|K(x)|(ν12g∞)2∗Cε122∗||v||2∗−λb3τ3ντ3W,τ3||v||τ3=(12−ε12esssupx∈RN|K(x)V(x)|)ξ||v||2−(12−ε12esssupx∈RN|K(x)V(x)|+esssupx∈RN|K(x)|(ν12g∞)2∗Cε122∗)||v||2∗−λb3τ3ντ3W,τ3||v||τ3 $
|
(4.51) |
for any $ v\in X $, $ \lambda > 0 $.
Consider
$ l_1(t) = (\frac{1}{2}-\frac{\varepsilon_1}{2}\text{ess}\sup\limits_{x\in\mathbb{R}^N}\big|\frac{K(x)}{V(x)}\big|)\xi t^2-\big(\frac{1}{2} -\frac{\varepsilon_1}{2}\text{ess}\sup\limits_{x\in\mathbb{R}^N}\big|\frac{K(x)}{V(x)}\big|+\text{ess}\sup\limits_{x\in\mathbb{R}^N}|K(x)|(\nu_1{\frac{2}{g_\infty}})^{2^*}\frac{C_{\varepsilon_1}}{22^*}\big)t^{2^*} $ |
for $ t\geq0 $. Obviously, there exists $ r_1 > 0 $ such that $ \max\limits_{t\geq0}l_1(t) = l_1(r_1)\triangleq \Lambda_1 > 0 $. Then, it follows from (4.51) that
$ Jλ(v)≥Λ1−λb3τντ3W,τ3rτ31 for any v∈Sr1,λ>0. $
|
(4.52) |
Set
$ \lambda_1 = \frac{\Lambda_1}{\frac{2b_3}{{\tau_3}}\nu_{W, {\tau_3}}^{\tau_3} r_1^{\tau_3}} $ |
and $ \alpha_1 = \frac{\Lambda_1}{2} $, and it follows from (4.52) that
$ Jλ(v)≥Λ1−λ1b3τ3ντ3W,τ3rτ31≥α1>0 for any λ∈(0,λ1), v∈Sr1. $
|
On the other hand, (4.51) implies that $ J_{\lambda}(v) $ is bounded blow in $ \mathcal{B}_{r_1} $ for any $ \lambda > 0 $. Taking $ \varphi\in X $ and $ \varphi\neq 0 $, by $ (g_{4}) $, $ (H'_{2}) $ and $ (W_{0}) $ and combining the Lebesgue dominated convergence theorem and Sobolev embedding theorem we have
$ limt→0+∫RNW(x)H(G−1(tφ))dx|t|τ3=limt→0+∫RNW(x)H(G−1(tφ))|G−1(tφ)|τ3|G−1(tφ)|τ3|tφ|τ3|φ|τ3dx=˜C∫RNW(x)|φ|τ3dx>0. $
|
Then, there exists $ \delta > 0 $ for any $ 0 < t < \delta $,
$ ∫RNW(x)H(G−1(tφ))dx>˜C∫RNW(x)|φ|τ3dx2tτ3. $
|
(4.53) |
Then, by $ (F'_{0}) $, $ K(x) > 0 $ for all $ x\in\mathbb{R}^N $, $ (g_{3}) $ and (4.53), we obtain that
$ Jλ(tφ)≤12∫RN|∇(tφ)|2dx+12∫RNV(x)|G−1(tφ)|2dx−λ∫RNW(x)H(G−1(tφ))dx≤t22(∫RN|∇φ|2dx+∫RNV(x)|φ|2dx)−λtτ3˜C∫RNW(x)|φ|τ3dx2 $
|
(4.54) |
for any $ 0 < t < \delta $, $ \lambda > 0 $. Since $ \tau_3\in(1, 2) $, there exists small $ t > 0 $ such that $ t\varphi\in \mathcal{B}_{r_1} $ and $ J_{\lambda}(t\varphi) < 0 $ for any $ \lambda > 0 $. Therefore, we complete the proof of this lemma.
Proof of Proposition 1.2. By Lemma 4.4 and Ekeland's variational principle [12], we infer that, for any $ \lambda\in(0, \lambda_1) $, there is a minimizing sequence $ \{v_n\}\subset\bar{\mathcal{B}}_{r_1} $ of the infimum $ c_0 = \inf\limits_{v\in\bar{\mathcal{B}}_{r_1}}J_{\lambda}(v) < 0 $, such that
$ c0≤Jλ(vn)≤c0+1n $
|
(4.55) |
and
$ Jλ(φ)≥Jλ(vn)−1n||φ−vn||, for all φ∈ˉBr1. $
|
(4.56) |
First, we claim that $ ||v_n|| < r_1 $ for large $ n\in\mathbb{N} $. Otherwise, we may assume that $ ||v_n|| = r_1 $. Up to a subsequence, by Lemma 4.4 we get $ J_{\lambda}(v_n)\geq\alpha_1 > 0 $, which and (4.55) imply that $ 0 > c_0\geq\alpha_1 > 0 $, which is a contradiction. In general, we suppose that $ ||v_n|| < r_1 $ for all $ n\in\mathbb{N} $. Next, we will show that $ J'_{\lambda}(v_n)\rightarrow 0 $ in $ X^* $. For any $ n\in\mathbb{N} $ and $ \varphi\in X $ with $ ||\varphi|| = 1 $, we choose sufficiently small $ \delta_n > 0 $ such that $ ||v_n+t\varphi|| < r_1 $ for all $ 0 < t < \delta_n $. It follows from (4.56) that
$ Jλ(vn+tφ)−Jλ(vn)t≥−1n. $
|
(4.57) |
Letting $ t\rightarrow 0^+ $, we get
$ \langle J'_{\lambda}(v_n), \varphi\rangle\geq-\frac{1}{n} $ |
for any $ n\in\mathbb{N} $. Similarly, replacing $ \varphi $ with $ -\varphi $ in the above arguments, we have
$ \langle J'_{\lambda}(v_n), \varphi\rangle\leq\frac{1}{n} $ |
for any $ n\in\mathbb{N} $. Therefore, we conclude that, for all $ \varphi\in X $ with
$ ||\varphi|| = 1, \ \ \langle J'_{\lambda}(v_n), \varphi\rangle\rightarrow 0\; \text{as}\; n\rightarrow \infty. $ |
Thus, we obtain that $ J_{\lambda}(v_n)\rightarrow c_0 $ and $ J'_{\lambda}(v_n)\rightarrow 0 $ as $ n\rightarrow \infty $. Noticing that $ ||v_n|| < r_1 $, we get that $ \{v_n\} $ is a $ (Ce)_{c_0} $ sequence for $ J_{\lambda}(v) $ in $ X $, and there exists $ v_*\in\bar{\mathcal{B}}_{r_1} $ such that $ v_n\rightharpoonup v_* $ in $ X $. Using the same type of arguments in Lemmas 4.1–4.3, we obtain that $ v_* $ is a critical point for $ J_{\lambda}(v) $ in $ X $ satisfying $ v_*\neq0 $ and $ J_{\lambda}(v_*) = c_0 < 0 $.
Next, we investigate the existence of ground state solutions for (1.1). For any $ \lambda\in(0, \lambda_1) $, define
$ \mathcal{S} = \{v\in X: J'_{\lambda}(v) = 0, \ v\neq 0\}\ \ \text{and}\ \ \mathcal{M}_0 = \inf\limits_{v\in\mathcal{S}}J_{\lambda}(v). $ |
Clearly, $ \mathcal{S} $ is nonempty and $ \mathcal{M}_0 < 0 $. For all $ v\in\mathcal{S} $, set
$ \varpi = G^{-1}(v)g(G^{-1}(v)). $ |
Then, we deduce from (1.7), $ K(x) > 0 $ for all $ x\in\mathbb{R}^N $, $ (F_{4}) $, $ (H'_{0}) $, $ (H'_{1}) $, $ (g_{2}) $, $ (W_{0}) $–$ (W_{2}) $ and Lemma 2.3 that for any $ \lambda\in(0, \lambda_1) $,
$Jλ(v)=Jλ(v)−1μ⟨J′λ(v),ϖ⟩≥(12−1+βμ)∫RN|∇v|2dx+(12−1μ)∫RNV(x)[G−1(v)]2dx−λb3τ3∫RNW(x)|G−1(v)|τ3dx≥(12−1+βμ)∫RN|∇G−1(v)|2dx+(12−1μ)∫RNV(x)[G−1(v)]2dx−λb3τ3∫RNW(x)|G−1(v)|τ3dx≥(12−1+βμ)||G−1(v)||2−λb3τ3ντ3W,τ3||G−1(v)||τ3. $
|
(4.58) |
Consider the function
$ l_2(t) = \big(\frac{1}{2}-\frac{1+\beta}{\mu}\big)t^2-\lambda\frac{b_3}{\tau_3}\nu_{W, \tau_3}^{\tau_3} t^{\tau_3} $ |
for $ t\geq 0 $. Since $ \tau_3\in (1, 2) $, for any fixed $ \lambda\in(0, \lambda_1) $ there exists $ t_2 > 0 $ such that
$ -\infty < \min\limits_{t\geq 0}l_2(t) = l_2(t_2) < 0. $ |
Then, it follows from (4.58) that $ J_{\lambda}(v)\geq l_2(t_2) > -\infty $ for any $ v\in\mathcal{S} $, which implies $ \mathcal{M}_0 > -\infty $. Letting $ \{v_n\}\subset \mathcal{S} $ be a minimizing sequence of $ \mathcal{M}_0 $ such that $ J_{\lambda}(v_n)\rightarrow \mathcal{M}_0 $, set
$ \varpi_n = G^{-1}(v_n)g(G^{-1}(v_n)). $ |
Since $ v_n\in\mathcal{S} $ for any $ n\in\mathbb{N} $, then $ \langle J'_{\lambda}(v_n), \varpi_n\rangle = 0 $ for any $ n\in\mathbb{N} $. Repeating the ideas explored in the proof of Lemma 3.3, we have that $ \{v_n\} $ is bounded in $ X $. Thus, $ \{v_n\} $ is a $ (Ce)_{\mathcal{M}_0} $ sequence and there exists $ v^*\in X $ such that $ v_n\rightharpoonup v^* $ in $ X $. Using the same type of arguments in Lemmas 4.1–4.3, we obtain that $ v^* $ is a critical point for $ J_{\lambda}(v) $ satisfying $ v^*\neq 0 $ and $ J_{\lambda}(v^*) = \mathcal{M}_0 < 0 $. Thus, $ v^* $ is a ground state solution of (2.4). This ends the proof of Proposition 1.2.
By using the variational method, this paper studies a kind of generalized quasilinear Schrodinger equation with concave-convex nonlinearities and potentials vanishing at infinity. We use the mountain pass theorem to prove that this problem has a positive solution. In addition, the existence of a ground state solution is also proved by Ekeland's variational principle. To the best of our knowledge, few works in this case seem to have appeared in the literature.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
We are grateful to the referees for their important suggestions, particularly for the suggestion in the ground state solution.
All authors declare no conflicts of interest in this paper.
[1] | Ruffman T, Henry JD, Livingstone V, et al. (2008) A meta-analytic review of emotion recognition and aging: Implications for neuropsychological models of aging. Neurosci Biobehav Rev 32: 863-881. |
[2] | Klein-Koerkamp Y, Beaudoin M, Baciu M, et al. (2012) Emotional decoding abilities in Alzheimer's disease: A meta-analysis. J Alzheimer's Disease 32: 109-125. |
[3] | Phillips LH, Scott C, Henry JD, et al. (2010) Emotion perception in Alzheimer's disease and mood disorder in old age. Psycholo Aging 25: 38-47. |
[4] | Henry JD, Ruffman T, McDonald S, et al. (2008) Recognition of disgust is selectively preserved in Alzheimer's disease. Neuropsycholo 46: 203-208. |
[5] |
Isaacowitz DM, Löckenhoff CE, Lane RD, et al. (2007) Age differences in recognition of emotion in lexical stimuli and facial expressions. Psycholo Aging 22: 147-159. doi: 10.1037/0882-7974.22.1.147
![]() |
[6] | de Gelder B, Van den Stock J (2011) The Bodily Expressive Action Stimulus Test (BEAST). Construction and Validation of a Stimulus Basis for Measuring Perception of Whole Body Expression of Emotions. Frontiers Psycholo 2: 181. |
[7] | de Gelder B, Van den Stock J, Meeren H, et al. (2010) Standing up for the body. Recent progress in uncovering the networks involved in the perception of bodies and bodily expressions. Neurosci Biobehav Rev 34 (4): 513-527. |
[8] | Johansson G (1973) Visual perception of biological motion and a model for its analysis. Perception Psychophysics 14: 201-211. |
[9] | Atkinson AP, Dittrich WH, Gemmell AJ, et al. (2004) Emotion perception from dynamic and static body expressions in point-light and full-light displays. Perception 33: 717-746. |
[10] | Heberlein AS, Adolphs R, Tranel D, et al. (2004) Cortical regions for judgments of emotions and personality traits from point-light walkers. J Cognitive Neurosci 16: 1143-1158. |
[11] | Ruffman T, Sullivan S, Dittrich W (2009) Older adults' recognition of bodily and auditory expressions of emotion. Psycholo Aging 24: 614-622. |
[12] | Calder AJ, Keane J, Manly T, et al. (2003) Facial expression recognition across the adult life span. Neuropsychologia 41: 195-202. |
[13] |
Insch PM, Bull R, Phillips LH, et al. (2012) Adult aging, processing style, and the perception of biological motion. Exper Aging Res 38: 169-185. doi: 10.1080/0361073X.2012.660030
![]() |
[14] | Billino J, Bremmer F, Gegenfurtner KR (2008) Differential aging of motion processing mechanisms: Evidence against general perceptual decline. Vision Res 48: 1254-1261. |
[15] | Pilz KS, Bennett PJ, Sekuler AB (2010) Effects of aging on biological motion discrimination. Vision Res 50: 211-219. |
[16] | Rosen HJ, Wilson MR, Schauer GF, et al. (2006) Neuroanatomical correlates of impaired recognition of emotion in dementia. Neuropsychologia 44: 365-373. |
[17] | Gilmore GC, Wenk HE, Naylor LA, et al. (1994) Motion perception and Alzheimers disease. J Gerontolo 49: 52-57. |
[18] | Sauer J, Ffytche DH, Ballard C, et al. (2006) Differences between Alzheimer's disease and dementia with Lewy bodies: an fMRI study of task-related brain activity. Brain 129: 1780-1788 |
[19] | Koff E, Zaitchik D, Montepare J, et al. (1999) Processing of emotion through the visual and auditory domains by patients with Alzheimer's disease. J Int Neuropsycholo Soc 5: 32-40. |
[20] |
Henry JD, Thompson C, Rendell PG, et al. (2012) Perception of biological motion and emotion in mild cognitive impairment and dementia. J Int Neuropsycholo Soc 18: 866-873. doi: 10.1017/S1355617712000665
![]() |
[21] | Sasson NJ, Pinkham AE, Richard J, et al. (2010) Controlling for response biases clarifies sex and age differences in facial affect recognition. J Nonverbal Behav 34: 207-221. |
[22] | Kumfor F, Piguet O (2012) Disturbance of emotion processing in frontotemporal dementia: a synthesis of cognitive and neuroimaging findings. Neuropsycholo Rev 22: 280-297 |
[23] | McKhann G, Drachman D, Folstein M (1984) Clinical diagnosis of Alzheimer's disease: report of the NINCDS-ADRDA work group under the auspices of department of health and human services task force on Alzheimer's disease. Neurolo 34: 939-944. |
[24] | Folstein MF, Folstein SE, McHugh PR (1975) 'Mini mental state'. A practical method for grading the cognitive state of patients for the clinician. J Psychiatric Res 12: 189-198. |
[25] | de Gelder B, Van den Stock J, Balaguer R, et al. (2008) Huntington's disease impairs recognition of angry and instrumental body language. Neuropsychologia 46 (1): 369-373. |
[26] | Van Hoesen GW, Parvizi J, Chu C (2000) Orbitofrontal cortex pathology in Alzheimer's disease. Cerebral Cortex 10: 243-251. |
[27] | Jones BF, Barnes J, Uylings HB, et al. (2006) Differential regional atrophy of the cingulate gyrus in Alzheimer disease: a volumetric MRI study. Cereb Cortex 16:1701-1708. |
[28] | Poulin SP, Dautoff R, Morris JC, et al. (2011) Amygdala atrophy is prominent in early Alzheimer's disease and relates to symptom severity. Psychiatry Res Neuroimaging 194: 7-13. |
[29] | Staff RT, Ahearn TS, Phillips LH, et al. (2011) The cerebral blood flow correlates of emotional facial processing in mild Alzheimer's disease. Neurosci Med 2: 6-13. |
[30] | Bucks RS, Radford SA (2004) Emotion processing in Alzheimer's disease. Aging Mental Health 8: 222-232. |
[31] | Cadieux NL, Greve KW (1997) Emotion processing in Alzheimer's disease. J Int Neuropsycholo Soc 3: 411-419. |
[32] |
Kohler CG, Anselmo-Gallagher G, Bilker W, et al. (2005) Emotion discrimination deficits in mild Alzheimer's disease. Am J Geriatric Psychiatry 13: 926-933. doi: 10.1097/00019442-200511000-00002
![]() |
[33] | Lavenu I, Pasquier F, Lebert F, et al. (1999) Perception of emotion in frontotemporal dementia and Alzheimer's disease. Alzheimer Disease Associated Disorders: 96-101. |
[34] | McDonald S, Flanagan S, Martin I, et al. (2004) The ecological validity of TASIT: a test of social perception. Neuropsycholo Rehabilitation 14: 285-302. |
[35] | Cooper CL, Phillips LH, Johnston M, et al. (2014) Links between emotion perception and social participation restriction following stroke. Brain Injury 28: 122-126. |
1. | Raúl M. Falcón, M. Venkatachalam, S. Gowri, G. Nandini, On the r-dynamic coloring of some fan graph families, 2021, 29, 1844-0835, 151, 10.2478/auom-2021-0039 | |
2. | T. Deepa, M. Venkatachalam, Ismail Naci Cangul, On r-dynamic chromatic number of some brick product graphs C(2n, 1, p), 2022, 15, 1793-5571, 10.1142/S1793557122501029 | |
3. | Raúl M. Falcón, M. Venkatachalam, S. Gowri, G. Nandini, On the r-dynamic coloring of the direct product of a path and a k-subdivision of a star graph, 2022, 14, 1793-8309, 10.1142/S1793830921501214 | |
4. | Raúl M. Falcón, Venkitachalam Aparna, Nagaraj Mohanapriya, Optimal secret share distribution in degree splitting communication networks, 2023, 18, 1556-1801, 1713, 10.3934/nhm.2023075 | |
5. | Raúl M. Falcón, S. Gowri, M. Venkatachalam, Solving the dynamic coloring problem for direct products of paths with fan graphs, 2023, 31, 1844-0835, 115, 10.2478/auom-2023-0006 |