1 Introduction

In this paper, we consider classification and criteria of the limit cases for the following singular second-order linear equation:

( p ( t ) y Δ ( t ) ) Δ +q(t) y σ (t)=λw(t) y σ (t),t[ρ(0),)T,
(1.1)

where p Δ , q, and w are real and piecewise continuous functions on [ρ(0),)T, p(t)0 and w(t)>0 for all t[ρ(0),)T; λC is the spectral parameter; T is a time scale with ρ(0)T and supT=; σ(t) and ρ(t) are the forward and backward jump operators in T; y Δ is the Δ-derivative of y; and y σ (t):=y(σ(t)).

The spectral problems of symmetric linear differential operators and difference operators can both be divided into two cases. Those defined over finite closed intervals with well-behaved coefficients are called regular. Otherwise, they are called singular. In 1910, Weyl [1] gave a dichotomy of the limit-point and limit-circle cases for the following singular second-order linear differential equation:

y (t)+q(t)y(t)=λy(t),t[0,),
(1.2)

where q is a real and continuous function on [0,), λC is the spectral parameter. Later, Titchmarsh, Coddington, Levinson etc. developed his results and established the Weyl-Titchmarsh theory [2, 3]. Their work has been greatly developed and generalized to higher-order differential equations and Hamiltonian systems, and a classification and some criteria of limit cases were formulated [49]. Singular spectral problems of self-adjoint scalar second-order difference equations over infinite intervals were firstly studied by Atkinson [10]. His work was followed by Hinton, Jirari etc. [11, 12]. In 2001, some sufficient and necessary conditions and several criteria of the limit-point and limit-circle cases were obtained for the following formally self-adjoint second-order linear difference equations with real coefficients [13]:

( p ( n ) Δ y ( n ) ) +q(n)y(n)=λw(n)y(n),n { n } n = 0 ,
(1.3)

where ∇ and Δ are the backward and forward difference operators respectively, namely y(n):=y(n)y(n1) and Δy(n):=y(n+1)y(n); p(n), q(n), and w(n) are real numbers with w(n)>0 for n[0,) and p(n)0 for n[1,); λ is a complex spectral parameter. In 2006, Shi [14] established the Weyl-Titchmarsh theory of discrete linear Hamiltonian systems. Later, several sufficient conditions and sufficient and necessary conditions for the limit-point and limit-circle cases were established for the singular second-order linear difference equation with complex coefficients (see [15]).

In the past twenty years, a lot of effort has been put into the study of regular spectral problems on time scales (see [1623]). But singular spectral problems have started to be considered only quite recently. In 2007, we employed Weyl’s method to divide the following singular second-order linear equations on time scales into two cases: limit-point and limit-circle cases [24]:

y Δ Δ (t)+q(t) y σ (t)=λ y σ (t),t[ρ(0),)T,

where q is real and continuous on [ρ(0),)T, λC is the spectral parameter. By using the similar method, Huseynov [25] studied the classification of limit cases for the following singular second-order linear equations on time scales:

( p ( t ) y Δ ( t ) ) +q(t)y(t)=λy(t),t(a,)T,

as well as of the form

( p ( t ) y Δ ( t ) ) Δ +q(t) y σ (t)=λ y σ (t),t[a,)T,
(1.4)

where p (or p Δ ) and q are real and piecewise continuous functions in (a,)T (or [a,)T), p(t)0 for all t, and λC is the spectral parameter. Obviously, let w(t)1 and ρ(0)=a, then (1.1) is the same as (1.4). In 2010, by using the properties of the Weyl matrix disks, Sun [26] established the Weyl-Titchmarsh theory of Hamiltonian systems on time scales. It has been found that the second-order singular differential and difference equations can be divided into limit-point and limit-circle cases. We wonder whether the classification of the limit cases holds on time scales. In the present paper, we extend these results obtained in [24] to Eq. (1.1) and establish several sufficient conditions and sufficient and necessary conditions for the limit-point and limit-circle cases for Eq. (1.1).

This paper is organized as follows. In Section 2, some basic concepts and a fundamental theory about time scales are introduced. In Section 3, a family of nested circles which converge to a limiting set is constructed. The dichotomy of the limit-point and limit-circle cases for singular second-order linear equations on time scales is given by the geometric properties of the limiting set. Finally, several criteria of the limit-point case are established, and the invariance of the limit cases is shown under a bounded perturbation for the potential function q in Section 4.

2 Preliminaries

In this section, some basic concepts and fundamental results on time scales are introduced.

Let TR be a non-empty closed set. The forward and backward jump operators σ,ρ:TT are defined by

σ(t):=inf{sT:s>t},ρ(t):=sup{sT:s<t},

respectively, where inf=supT, sup=infT. A point tT is called right-scattered, right-dense, left-scattered, and left-dense if σ(t)>t,σ(t)=t,ρ(t)<t, and ρ(t)=t separately. Denote T k :=T if T is unbounded above and T k :=T(ρ(maxT),maxT] otherwise. The graininess μ:T[0,) is defined by

μ(t):=σ(t)t.

Let f be a function defined on T. f is said to be Δ-differentiable at t T k provided there exists a constant a such that, for any ε>0, there is a neighborhood U of t (i.e., U=(tδ,t+δ)T for some δ>0) with

| f ( σ ( t ) ) f ( s ) a ( σ ( t ) s ) | ε | σ ( t ) s | for all sU.

In this case, denote f Δ (t):=a. If f is Δ-differentiable for every t T k , then f is said to be Δ-differentiable on T. If f is Δ-differentiable at t T k , then

f Δ (t)={ lim s T s t f ( t ) f ( s ) t s , if  μ ( t ) = 0 , f ( σ ( t ) ) f ( t ) μ ( t ) , if  μ ( t ) > 0 .
(2.1)

If F Δ (t)=f(t) for all t T k , then F(t) is called an anti-derivative of f on T. In this case, define the Δ-integral by

s t f(τ)Δτ=F(t)F(s)for all s,tT.

For convenience, we introduce the following results ([[27], Chapter 1] and [[28], Chapter 1]), which are useful in this paper.

Lemma 2.1 Let f,g:TR and t T k .

  1. (i)

    If f is Δ-differentiable at t, then f is continuous at t.

  2. (ii)

    If f and g are Δ-differentiable at t, then fg is Δ-differentiable at t and

    ( f g ) Δ (t)= f σ (t) g Δ (t)+ f Δ (t)g(t)= f Δ (t) g σ (t)+f(t) g Δ (t).
  3. (iii)

    If f and g are Δ-differentiable at t, and f(t) f σ (t)0, then f 1 g is Δ-differentiable at t and

    ( g f 1 ) Δ (t)= ( g Δ ( t ) f ( t ) g ( t ) f Δ ( t ) ) ( f σ ( t ) f ( t ) ) 1 .

A function f defined on T is said to be rd-continuous if it is continuous at every right-dense point in T and its left-sided limit exists at every left-dense point in T. The set of rd-continuous functions f:TR is denoted by C r d (T)= C r d (T,R). The set of k th Δ-differentiable functions with rd-continuous k th derivative is denoted by C r d k (T)= C r d k (T,R).

Lemma 2.2 If f, g are rd-continuous functions on T, then

  1. (i)

    f σ is rd-continuous and f has an anti-derivative on T;

  2. (ii)

    t σ ( t ) f(τ)Δτ=μ(t)f(t) for all tT.

  3. (iii)

    (Integration by parts) a b f σ (τ) g Δ (τ)Δτ=f(b)g(b)f(a)g(a) a b f Δ (τ)g(τ)Δτ.

  4. (iv)

    (Hölder’s inequality [[29], Lemma 2.2(iv)]) Let r,sT with rs, then

    r s | f ( τ ) g ( τ ) | Δτ { r s | f ( τ ) | p Δ τ } 1 p { r s | g ( τ ) | q Δ τ } 1 q ,

where p>1 and q=p/(p1).

Let

L w 2 ( ρ ( 0 ) , ) := { y σ : [ ρ ( 0 ) , ) C | ρ ( 0 ) w ( t ) | y σ ( t ) | 2 Δ t < } .

A function g:TR is called regressive if

1+μ(t)g(t)0for all tT.

Higer [30] showed that for any given t 0 T and for any given rd-continuous and regressive g, the initial value problem

y Δ (t)=g(t)y(t),y( t 0 )=1

has a unique solution

e g ( t , t 0 ) = exp { t 0 t ξ μ ( τ ) ( g ( τ ) ) Δ τ } , ξ h ( z ) = { Log ( 1 + h z ) h , if  h 0 , z , if  h = 0 .
(2.2)

Lemma 2.3 ([[27], Theorem 6.1])

Let y,f C r d (T) and g R + :={g C r d (T):1+μ(t)g(t)>0,tT}. Then

y Δ (t)g(t)y(t)+f(t),tT,

implies

y(t)y( t 0 ) e g (t, t 0 )+ t 0 t e g ( t , σ ( τ ) ) f(τ)Δτ,tT.

We define the Wronskian by

W[x,y](t)=p(t) [ x ( t ) y Δ ( t ) x Δ ( t ) y ( t ) ] ,x,y C r d 2 (T).
(2.3)

The following result is a direct consequence of the Lagrange identity [[27], Theorem 4.30].

Lemma 2.4 Let x and y be any two solutions of (1.1). Then W[x,y](t) is a constant in [ρ(0),)T.

3 Classification

In this section, we focus on the classification of the limit cases for singular second-order linear equations on time scales.

Let y 1 (t,λ) and y 2 (t,λ) be the two solutions of (1.1) satisfying the following initial conditions:

respectively. Since their Wronskian is identically equal to 1, these two solutions form a fundamental solution system of (1.1). We form a linear combination of y 1 (t,λ) and y 2 (t,λ)

y(t,λ,m):= y 1 (t,λ)+m y 2 (t,λ).
(3.1)

Let b(ρ(0),)T, kR, λ=μ+iν with ν0, and let (3.1) satisfy

p(b) y Δ (b,λ,m)+ky(b,λ,m)=0.
(3.2)

Then

m= p ( b ) y 1 Δ ( b , λ ) + k y 1 ( b , λ ) p ( b ) y 2 Δ ( b , λ ) + k y 2 ( b , λ ) .
(3.3)

It can be verified that the integral identity

[ y ( t , λ ) ¯ p ( t ) y Δ ( t , λ ) ] | t 1 t 2 t 1 t 2 p(t) | y Δ ( t , λ ) | 2 Δt+ t 1 t 2 [ λ w ( t ) q ( t ) ] | y σ ( t , λ ) | 2 Δt=0
(3.4)

holds for any solution y(t,λ) of (1.1) and for any t 1 , t 2 [ρ(0),)T. Setting y(t,λ)= y 2 (t,λ), t 1 =ρ(0), t 2 =b in (3.4) and taking its imaginary part, we obtain

[ y 2 ( b , λ ) ¯ p ( b ) y 2 Δ ( b , λ ) ] =ν ρ ( 0 ) b w(t) | y 2 σ ( t , λ ) | 2 Δt.
(3.5)

So

( p ( b ) y 2 Δ ( b , λ ) y 2 ( b , λ ) ) = [ y 2 ( b , λ ) ¯ p ( b ) y 2 Δ ( b , λ ) ] | y 2 ( b , λ ) | 2 0.

It follows from (3.2) and kR that k p ( b ) y 2 Δ ( b , λ ) y 2 ( b , λ ) . Hence, the denominator in (3.3) is not equal to zero, and consequently, m is well defined.

Next, we will show that (3.3) describes a circle for any fixed b. It follows from (3.1) and (3.2) that

y ( ρ ( 0 ) , λ , m ) ¯ p ( ρ ( 0 ) ) y Δ ( ρ ( 0 ) , λ , m ) =m

and

y ( b , λ , m ) ¯ p(b) y Δ (b,λ,m)=k | y ( b , λ , m ) | 2 R.

By (3.4) and the above two relations, we have

(m)=ν ρ ( 0 ) b w(t) | y σ ( t , λ , m ) | 2 Δt,
(3.6)

which implies that m lies in the upper half-plane if ν>0. It follows from (3.2) that

k= p ( b ) y Δ ( b , λ , m ) y ( b , λ , m ) ,

which, together with kR, yields that

p(b) [ y Δ ( b , λ , m ) y ( b , λ , m ) ¯ y Δ ( b , λ , m ) ¯ y ( b , λ , m ) ] =0.

It is equivalent to

W[y, y ¯ ](b,λ,m)=0.
(3.7)

By using (3.1), (3.7) can be expanded as

| m | 2 W[ y 2 , y 2 ¯ ](b,λ)+mW[ y 2 , y 1 ¯ ](b,λ)+ m ¯ W[ y 1 , y 2 ¯ ](b,λ)+W[ y 1 , y 1 ¯ ](b,λ)=0.
(3.8)

Moreover, setting m=u+iv, we have

W[ y 2 , y 2 ¯ ](b,λ)=2iA,W[ y 1 , y 1 ¯ ](b,λ)=2iD,W[ y 2 , y 1 ¯ ](b,λ)=B+iC.
(3.9)

It follows from the last relation in (3.9) that we have W[ y 1 , y 2 ¯ ](b,λ)=BiC. By using (2.3) and (3.9), it can be verified that

(3.10)

It follows from the first relation in (3.9) and (3.5) that we have A=ν ρ ( 0 ) b w(t) | y 2 σ ( t , λ ) | 2 Δt0. Then (3.8) becomes

( u C 2 A ) 2 + ( v B 2 A ) 2 = B 2 + C 2 4 A D 4 A 2 ,
(3.11)

which implies that (3.3) forms a circle C b as k varies. It is evident that the center of C b is

z 0 = C + i B 2 A = B i C 2 i A = W [ y 1 , y 2 ¯ ] ( b , λ ) W [ y 2 , y 2 ¯ ] ( b , λ ) .

It follows from Lemma 2.4 and (3.10) that

B 2 + C 2 4AD= | W [ y 1 , y 2 ] ( b , λ ) | 2 = | W [ y 1 , y 2 ] ( ρ ( 0 ) , λ ) | 2 =1.

From (3.11), (3.9), (2.3), and (3.5) we have that the radius of C b is

r b = | B 2 + C 2 4 A D 4 A 2 | 1 2 = | 2 i A | 1 = | W [ y 2 , y 2 ¯ ] ( b , λ ) | 1 = [ 2 | ν | ρ ( 0 ) b w ( t ) | y 2 σ ( t , λ ) | 2 Δ t ] 1 .
(3.12)

Let C b ¯ denote the closed disk bounded by C b . We are going to show that the circle sequence { C b ¯ }(ρ(0)<b<) is nested.

Set

U+iV=ν ρ ( 0 ) b w(t) y 1 σ (t,λ) y 2 σ ( t , λ ) ¯ Δt.

From the first relation in (3.9), we have

A=ν ρ ( 0 ) b w(t) | y 2 σ ( t , λ ) | 2 Δt.

Similarly,

D=ν ρ ( 0 ) b w(t) | y 1 σ ( t , λ ) | 2 Δt.

So, it follows from (3.6) that

v=A ( u 2 + v 2 ) +2Uu+2Vv+D.
(3.13)

In the case of ν>0, the point m=u+iv is interior to the circle if v>A( u 2 + v 2 )+2Uu+2Vv+D. This shows that m C b ¯ if and only if

(m)ν ρ ( 0 ) b w(t) | y σ ( t , λ , m ) | 2 Δt.

Let b 1 , b 2 [ρ(0),)T with b 1 < b 2 and consider the corresponding disks C b 1 ¯ and C b 2 ¯ . For any m C b 2 ¯ , we have

(m)ν ρ ( 0 ) b 2 w(t) | y σ ( t , λ , m ) | 2 Δtν ρ ( 0 ) b 1 w(t) | y σ ( t , λ , m ) | 2 Δt.

Hence, m C b 1 ¯ . This yields that C b 2 ¯ C b 1 ¯ . Therefore, { C b ¯ } is nested. Consequently, there are the following two alternatives:

  1. (1)

    r b 0 as b. In this case there is one point m=m(λ) which is common to all the disks C b ¯ , b[ρ(0),)T. This is called the limit-point case. It follows from (3.12) that this case occurs if and only if

    ρ ( 0 ) w(t) | y 2 σ ( t , λ ) | 2 Δt=.
    (3.14)
  2. (2)

    r b r >0 as b. In this case there is a disk C ¯ contained in all the disks C b ¯ , b[ρ(0),)T. This is called the limit-circle case. It follows from (3.12) that this case occurs if and only if the integral in (3.14) is convergent, i.e., y 2 (,λ) L w 2 (ρ(0),).

Theorem 3.1 For every non-real λC, Eq. (1.1) has at least one non-trivial solution in L w 2 (ρ(0),).

Proof In the limit-circle case, it follows from the above discussion that y 2 (,λ) L w 2 (ρ(0),).

Next, we will show that y 1 (,λ)+m(λ) y 2 (,λ) L w 2 (ρ(0),) in the limit-point case. Let { b n }T with 0< b n < b n + 1 and choose any m n C b n . Then m n m(λ) as n and y σ (t,λ, m n ) uniformly converges to y σ (t,λ,m(λ)) on any finite interval [ρ(0),ω]T, ωT. Since the sequence {( m n )} is bounded from above and its upper bound is denoted by y 0 , then for b n >ω,

y 0 ( m n )=ν ρ ( 0 ) b n w(t) | y σ ( t , λ , m n ) | 2 Δtν ρ ( 0 ) ω w(t) | y σ ( t , λ , m n ) | 2 Δt.

Hence, by the uniform convergence of y σ (t,λ, m n ), we have

y 0 ν ρ ( 0 ) ω w(t) | y σ ( t , λ , m ( λ ) ) | 2 Δt

for all ω. Therefore, y(,λ,m(λ))= y 1 (,λ)+m(λ) y 2 (,λ) L w 2 (ρ(0),). This completes the proof. □

Remark 3.1 Similar to the proof of Theorem 3.1, it can be easily verified that y(,λ,m) L w 2 (ρ(0),) for any m C with (λ)0 in the limit-circle case. Clearly, y(t,λ,m) and y 2 (t,λ) are linearly independent. Hence, all the solutions of Eq. (1.1) belong to L w 2 (ρ(0),) for any λC with (λ)0 in the limit-circle case.

Remark 3.2 It follows from (3.14) and Theorem 3.1 that Eq. (1.1) has exactly one linearly independent solution in L w 2 (ρ(0),) in the limit point case for any λC with (λ)0.

Theorem 3.2 If Eq. (1.1) has two linearly independent solutions in L w 2 (ρ(0),) for some λ 0 C, then this property holds for all λC.

Proof Suppose that Eq. (1.1) has two linearly independent solutions in L w 2 (ρ(0),) for λ= λ 0 C. Then y 1 (t, λ 0 ) and y 2 (t, λ 0 ) are in L w 2 (ρ(0),). For briefness, denote

u 1 (t)= y 1 (t, λ 0 ), u 2 (t)= y 2 (t, λ 0 ).

For any λC, let v(t) be an arbitrary non-trivial solution of (1.1), and let u(t) be the solution of (1.1) with λ= λ 0 and with the initial values

u(a)=v(a), u Δ (a)= v Δ (a),a(0,)T.

From the variation of constants [[27], Theorem 3.73], we have

v(t)=u(t)+(λ λ 0 ) a t [ u 1 ( t ) u 2 σ ( s ) u 2 ( t ) u 1 σ ( s ) ] w(s) v σ (s)Δs,t[a,)T.
(3.15)

Replacing t with σ(t) in (3.15) and using (ii) of Lemma 2.2, we obtain

which implies by the Hölder inequality in Lemma 2.2 that

| w 1 2 ( t ) v σ ( t ) | | w 1 2 ( t ) u σ ( t ) | + | λ λ 0 | | w 1 2 ( t ) u 1 σ ( t ) | × [ a t w ( s ) | u 2 σ ( s ) | 2 Δ s a t w ( s ) | v σ ( s ) | 2 Δ s ] 1 2 + | λ λ 0 | | w 1 2 ( t ) u 2 σ ( t ) | [ a t w ( s ) | u 1 σ ( s ) | 2 Δ s a t w ( s ) | v σ ( s ) | 2 Δ s ] 1 2 .

It follows from the inequality

( A + B + C ) 2 3 ( A 2 + B 2 + C 2 ) ,

where A, B, C are non-negative numbers, that

1 3 w ( t ) | v σ ( t ) | 2 w ( t ) | u σ ( t ) | 2 + | λ λ 0 | 2 [ w ( t ) | u 1 σ ( t ) | 2 a t w ( s ) | u 2 σ ( s ) | 2 Δ s + w ( t ) | u 2 σ ( t ) | 2 a t w ( s ) | u 1 σ ( s ) | 2 Δ s ] a t w ( s ) | v σ ( s ) | 2 Δ s .

Integrating the two sides of the above inequality with respect to t from a to τ(a,)T, we get

1 3 a τ w ( t ) | v σ ( t ) | 2 Δ t a τ w ( t ) | u σ ( t ) | 2 Δ t + | λ λ 0 | 2 a τ [ w ( t ) | u 1 σ ( t ) | 2 a t w ( s ) | u 2 σ ( s ) | 2 Δ s + w ( t ) | u 2 σ ( t ) | 2 a t w ( s ) | u 1 σ ( s ) | 2 Δ s ] a t w ( s ) | v σ ( s ) | 2 Δ s Δ t ,

which yields that

Hence,

(3.16)

The constant a can be chosen in advance so large that

6 | λ λ 0 | 2 a w(t) | u 1 σ ( t ) | 2 Δt a w(t) | u 2 σ ( t ) | 2 Δt<1.

It follows from (3.16) that v L w 2 (a,) and hence v L w 2 (ρ(0),). Therefore, all the solutions of Eq. (1.1) are in L w 2 (ρ(0),). The proof is complete. □

At the end of this section, from the above discussions we present the classification of the limit cases for singular second-order linear equations over the infinite interval [ρ(0),)T on time scales.

Definition 3.1 If Eq. (1.1) has only one linear independent solution in L w 2 (ρ(0),) for some λC, then Eq. (1.1) is said to be in the limit-point case at t=. If Eq. (1.1) has two linear independent solutions in L w 2 (ρ(0),) for some λC, then Eq. (1.1) is said to be in the limit-circle case at t=.

4 Several criteria of the limit-point and limit-circle cases

In this section, we establish several criteria of the limit-point and limit-circle cases for Eq. (1.1).

We first give two criteria of the limit-point case.

Theorem 4.1 Let w(t)1 and p(t)>0 for all t[ρ(0),)T. If there exists a positive Δ-differentiable function M(t) on [a,)T for some aρ(0) and two positive constants k 1 and k 2 such that for all t[a,)T,

  1. (i)

    q(t) k 1 M σ (t),

  2. (ii)

    p 1 2 (t)| M Δ (t)| ( M ( t ) ) 1 ( M σ ( t ) ) 1 2 k 2 ,

  3. (iii)

    a ( p ( t ) M σ ( t ) ) 1 2 Δt=,

then Eq. (1.1) is in the limit-point case at t=.

Proof Suppose that Eq. (1.1) is in the limit-circle case at t=. By Theorem 3.2, all the solutions of

( p ( t ) y Δ ( t ) ) Δ +q(t) y σ (t)=0,t[ρ(0),)T
(4.1)

are in L w 2 (ρ(0),). Let y 1 (t) and y 2 (t) be the solutions of (4.1) satisfying the following initial conditions:

y 1 ( ρ ( 0 ) ) =p ( ρ ( 0 ) ) y 2 Δ ( ρ ( 0 ) ) =0,p ( ρ ( 0 ) ) y 1 Δ ( ρ ( 0 ) ) = y 2 ( ρ ( 0 ) ) =1.
(4.2)

It is evident that y 1 (t) and y 2 (t) are two linearly independent solutions of (4.1) in L w 2 (ρ(0),). By Lemma 2.4, W[ y 1 , y 2 ](t)1 for all t[ρ(0),)T. Hence, we have

It follows from the Hölder inequality and assumption (iii) that

a p ( τ ) ( y 1 Δ ( τ ) ) 2 M σ ( τ ) Δτor a p ( τ ) ( y 2 Δ ( τ ) ) 2 M σ ( τ ) Δτ

are divergent. Suppose

a p ( τ ) ( y 1 Δ ( τ ) ) 2 M σ ( τ ) Δτ=.

From (4.1) and assumption (i), we have

a t y 1 σ ( τ ) [ p ( τ ) y 1 Δ ( τ ) ] Δ M σ ( τ ) Δ τ = a t q ( τ ) ( y 1 σ ( τ ) ) 2 M σ ( τ ) Δ τ k 1 a t ( y 1 σ ( τ ) ) 2 Δ τ .
(4.3)

Applying integration by parts in Lemma 2.2, by (iii) in Lemma 2.1, we get

(4.4)

Again applying the Hölder inequality, from condition (ii), we have

(4.5)

where

H(t):= a t p ( τ ) ( y 1 Δ ( τ ) ) 2 M σ ( τ ) Δτ.

Since

a ( y 1 σ ( τ ) ) 2 Δτ> a t ( y 1 σ ( τ ) ) 2 Δτ,

it follows from (4.3)-(4.5) that

It follows from the assumption that H(t) as t. From the above relation and p(t)>0 for all t[ρ(0),)T, we have that y 1 (t) y 1 Δ (t) is ultimately positive. Therefore, y 1 (t)0 as t; and consequently, y 1 (t) does not belong to L w 2 (ρ(0),). This contradicts the assumption that all the solutions of (4.1) are in L w 2 (ρ(0),). Then Eq. (4.1) has at least one non-trivial solution outside of L w 2 (ρ(0),). It follows from Theorem 3.2 that Eq. (1.1) is in the limit-point case at t=. This completes the proof. □

Remark 4.1 Since R and N are two special time scales, Theorem 4.1 not only contains the criterion of the limit-point case for second-order differential equations [[5], Chapter 9, Theorem 2.4], but also the criterion of the limit-point case for second-order difference equation (1.3) [[15], Theorem 3.3].

The following corollary is a direct consequence of Theorem 4.1 by setting M(t)1 for t[ρ(0),)T.

Corollary 4.1 If w(t)1, p(t)>0, q(t) is bounded below in [ρ(0),)T, and ρ ( 0 ) ( p ( t ) ) 1 2 Δt=, then Eq. (1.1) is in the limit-point case at t=.

Theorem 4.2 If

ρ ( 0 ) μ σ ( t ) [ w ( t ) w σ ( t ) ] 1 2 | p σ ( t ) | Δt=,
(4.6)

then Eq. (1.1) is in the limit-point case at t=.

Proof On the contrary, suppose that Eq. (1.1) is in the limit-circle case at t=. Let y 1 (t) and y 2 (t) be two linearly independent solutions of (1.1) in L w 2 (ρ(0),) satisfying the initial conditions (4.2). By Lemma 2.4, we have

W[ y 1 , y 2 ](t)=W[ y 1 , y 2 ] ( ρ ( 0 ) ) 1,t[ρ(0),)T,

which, together with (2.1), implies that

So, we get

| y 1 ( t ) | | y 2 σ ( t ) | + | y 2 ( t ) | | y 1 σ ( t ) | μ ( t ) | p ( t ) | ,t[ρ(0),)T,

which implies

(4.7)

where y σ 2 (t)= y σ (σ(t)). By the Hölder inequality and the assumption that y 1 , y 2 L w 2 (ρ(0),), one has

Hence, it follows from (4.7) that

ρ ( 0 ) μ σ ( t ) [ w ( t ) w σ ( t ) ] 1 2 | p σ ( t ) | Δt<,

which is a contradiction to the assumption (4.6). Therefore, Eq. (1.1) is in the limit-point case at t=. This completes the proof. □

Remark 4.2 Let T=N, Theorem 4.2 is the same as that obtained by Chen and Shi for second-order difference equations [[13], Corollary 3.1].

Next, we study the invariance of the limit cases under a bounded perturbation for the potential function q. Let f(t)=M and p(t)=f(t) in [[27], Theorem 2.4(i)]. It follows from [[27], Theorem 2.36(i)], [[27], Theorem 2.39(i)], and [[27], Theorem 2.4(i)] that we have the following lemma, which is useful in the subsequent discussion.

Lemma 4.1 (Gronwall’s inequality)

Let y,f C r d (T) be two non-negative functions on [ρ(0),)T and M be a non-negative constant. If

y(t)M+ ρ ( 0 ) t f(τ)y(τ)Δτ for all t[ρ(0),)T,
(4.8)

then

y(t)M e f ( t , ρ ( 0 ) ) for all t[ρ(0),)T,

where e f (t,s) is defined as in (2.2).

The following result shows that if Eq. (1.1) is in the limit-circle case, so is it under a bounded perturbation for the potential function q.

Lemma 4.2 Let q(t)=d(t)+e(t) for all t[ρ(0),)T and e(t) be bounded with respect to w(t) on [ρ(0),)T; that is, there exists a positive constant M such that

| e ( t ) | Mw(t),t[ρ(0),)T.
(4.9)

Then Eq. (1.1) is in the limit-circle case at t= if and only if the equation

( p ( t ) y Δ ( t ) ) Δ +d(t) y σ (t)=λw(t) y σ (t)
(4.10)

is in the limit-circle case at t=.

Proof Suppose that (4.10) is in the limit-circle case at t=. To show that Eq. (1.1) is in the limit-circle case, it suffices to show that each solution (4.1) is in L w 2 (ρ(0),) by Theorem 3.2.

Let y 1 (t) and y 2 (t) be two solutions of the equation

( p ( t ) y Δ ( t ) ) Δ +d(t) y σ (t)=0
(4.11)

satisfying the initial conditions (4.2). Then y 1 (t), y 2 (t) are two linearly independent solutions in L w 2 (ρ(0),) by Theorem 3.2.

Let y(t) be any solution of (4.1). Then

( p ( t ) y Δ ( t ) ) Δ +d(t) y σ (t)=r(t)for all t[ρ(0),)T,

where r(t):=e(t) y σ (t). By the variation of constants [[27], Theorem 3.73] there exist two constants α and β such that

y(t)=α y 1 (t)+β y 2 (t)+ ρ ( 0 ) t r(τ) ( y 1 σ ( τ ) y 2 ( t ) y 2 σ ( τ ) y 1 ( t ) ) Δτfor all t[ρ(0),)T.

Hence, replacing t by σ(t) and by (ii) in Lemma 2.2, we get

y σ (t)=α y 1 σ (t)+β y 2 σ (t)+ ρ ( 0 ) t r(τ) ( y 1 σ ( τ ) y 2 σ ( t ) y 2 σ ( τ ) y 1 σ ( t ) ) Δτ.
(4.12)

From (4.9) and (4.12), we have

| y σ ( t ) | | α | | y 1 σ ( t ) | + | β | | y 2 σ ( t ) | + M ρ ( 0 ) t ( | y 1 σ ( τ ) | | y 2 σ ( t ) | + | y 2 σ ( τ ) | | y 1 σ ( t ) | ) w ( τ ) | y σ ( τ ) | Δ τ .
(4.13)

Since y 1 (t), y 2 (t) are solutions of Eq. (4.11), which satisfy the initial conditions (4.2), it follows from the existence-uniqueness theorem that | y 1 σ (t)|+| y 2 σ (t)|0 for all t[ρ(0),)T. Let

y 0 σ (t):= | y σ ( t ) | | y 1 σ ( t ) | + | y 2 σ ( t ) | for all t[ρ(0),)T.

From (4.13), we have

y 0 σ ( t ) | α | | y 1 σ ( t ) | + | β | | y 2 σ ( t ) | | y 1 σ ( t ) | + | y 2 σ ( t ) | + M ρ ( 0 ) t ( | y 1 σ ( τ ) | | y 2 σ ( t ) | + | y 2 σ ( τ ) | | y 1 σ ( t ) | ) w ( τ ) | y σ ( τ ) | | y 1 σ ( t ) | + | y 2 σ ( t ) | Δ τ | α | + | β | + M ρ ( 0 ) t ( | y 1 σ ( τ ) | + | y 2 σ ( τ ) | ) w ( τ ) | y σ ( τ ) | Δ τ = | α | + | β | + M ρ ( 0 ) t ( | y 1 σ ( τ ) | + | y 2 σ ( τ ) | ) 2 w ( τ ) | y 0 σ ( τ ) | Δ τ | α | + | β | + 2 M ρ ( 0 ) t ( | y 1 σ ( τ ) | 2 + | y 2 σ ( τ ) | 2 ) w ( τ ) | y 0 σ ( τ ) | Δ τ .

It follows from (i) of Lemma 2.2 that y 0 σ () C r d (T). By Lemma 4.1, we have

y 0 σ ( t ) ( | α | + | β | ) e [ 2 M w ( | y 1 σ | 2 + | y 2 σ | 2 ) ] ( t , ρ ( 0 ) ) = ( | α | + | β | ) exp [ ρ ( 0 ) t ξ μ ( τ ) ( 2 M w ( τ ) ( | y 1 σ ( τ ) | 2 + | y 2 σ ( τ ) | 2 ) ) Δ τ ] = ( | α | + | β | ) exp [ ρ ( 0 ) t 1 μ ( τ ) Log ( 1 + μ ( τ ) 2 M w ( τ ) ( | y 1 σ ( τ ) | 2 + | y 2 σ ( τ ) | 2 ) ) Δ τ ] ( | α | + | β | ) exp [ ρ ( 0 ) t 2 M w ( τ ) ( | y 1 σ ( τ ) | 2 + | y 2 σ ( τ ) | 2 ) Δ τ ] ( | α | + | β | ) exp [ ρ ( 0 ) 2 M w ( τ ) ( | y 1 σ ( τ ) | 2 + | y 2 σ ( τ ) | 2 ) Δ τ ] = : C < ,

which implies that | y σ (t)|C(| y 1 σ (t)|+| y 2 σ (t)|). Hence, y() L w 2 (ρ(0),); and consequently, Eq. (1.1) is in the limit-circle case at t=.

On the other hand, using

( p ( t ) y Δ ( t ) ) Δ +d(t) y σ (t)= ( p ( t ) y Δ ( t ) ) Δ + ( q ( t ) e ( t ) ) y σ (t)

one can easily conclude that if Eq. (1.1) is in the limit-circle case, then Eq. (4.10) is in the limit-circle case. This completes the proof. □

Theorem 4.3 Let q(t)=d(t)+e(t) for all t[ρ(0),)T and e(t) be bounded with respect to w(t) on [ρ(0),)T. Then the limit cases for Eq. (1.1) are invariant.

Remark 4.3 Lemma 4.2 extends the related result [[13], Lemma 2.4] for the singular second-order difference equation to the time scales. In addition, let T=R in Lemma 4.2, then we can directly prove [[31], Theorem 6.1] with the similar method.