Home Curvature conditions, Liouville-type theorems and Harnack inequalities for a nonlinear parabolic equation on smooth metric measure spaces
Article Open Access

Curvature conditions, Liouville-type theorems and Harnack inequalities for a nonlinear parabolic equation on smooth metric measure spaces

  • Ali Taheri EMAIL logo and Vahideh Vahidifar
Published/Copyright: April 11, 2024

Abstract

In this paper we prove gradient estimates of both elliptic and parabolic types, specifically, of Souplet-Zhang, Hamilton and Li-Yau types for positive smooth solutions to a class of nonlinear parabolic equations involving the Witten or drifting Laplacian on smooth metric measure spaces. These estimates are established under various curvature conditions and lower bounds on the generalised Bakry-Émery Ricci tensor and find utility in proving elliptic and parabolic Harnack-type inequalities as well as general Liouville-type and other global constancy results. Several applications and consequences are presented and discussed.

1991 Mathematics Subject Classification: 53C44; 58J60; 58J35; 60J60

1 Introduction

Unravelling the interplay between the analytic, stochastic and geometric properties of Riemannian manifolds on the one hand and the dynamics of the solutions to nonlinear evolution equations on the other lies at the heart of geometric analysis. Some particular questions include: short and long time existence, smoothing properties, convergence to equilibrium, spectral gaps, blow-ups, self-similar solutions and their profiles, analysis of singularities, critical exponents, description and classification of ancient and eternal solutions as well as Liouville or global constancy type results to mention a few.

The ubiquitous gradient estimates play a significant role here and their derivation is often the main step in confronting such challenging tasks. For parabolic and heat-type equations, there are several types of gradient estimates and techniques in literature, each having their own strength and utility. In this paper we extend such techniques and results to the context of nonlinear equations on smooth metric measure spaces where understanding the interaction between the nonlinearity and curvature on the one hand and the different terms in the estimate on the other is the main driving force (see [1]–[36] for more).

Towards this end let (M, g) be a smooth Riemannian manifold of dimension n ≥ 2 and let dμ = eϕ dv g denote the weighted measure associated with a potential ϕ and the Riemannian volume measure dv g on M. The triple (M, g, dμ) is referred to as a smooth metric space or a weighted manifold. In this paper we develop gradient estimates of elliptic and parabolic types, more specifically, of Souplet-Zhang, Hamilton and Li-Yau types, along with some of their implications, for positive smooth solutions w = w(x, t) to the nonlinear parabolic equation on (M, g, dμ) given by:

(1.1) ( P ) w t = Δ ϕ w + G ( t , x , w ) , ( x , t ) M × ( , ) .

The operator Δ ϕ appearing on the right in (1.1) is the Witten Laplacian associated with the smooth metric measure space (M, g, dμ) (also called the weighted or drifting or ϕ-Laplacian) whose action on functions v C 2 ( M ) is given by

(1.2) Δ ϕ v = e ϕ div ( e ϕ v ) = Δ v ϕ , v .

Here Δ, div and ∇ are the usual Laplace-Beltrami, divergence and gradient operators associated with the metric g. The nonlinearity G = G ( t , x , w ) is a sufficiently smooth function of the space-time variables (t, x) as well as the dependent variable w. The general form of the assumed nonlinearity here enables us to more clearly examine the form and extent in which it influences the various gradient estimates and subsequently the implication it has on the qualitative properties of solutions.

Gradient estimates for positive solutions to linear and nonlinear heat type equations have been studied extensively starting from the seminal paper of Li and Yau [15] (see also [14]). In the nonlinear setting perhaps the first equation to be considered is the one with a logarithmic type nonlinearity (see, e.g., [37]) [specifically, G = A ( x , t ) w log w in ( P ) ] given by

(1.3) w t = Δ ϕ w + A ( x , t ) w log w .

The interest in such problems originates partly from its natural links with gradient Ricci solitons and partly from links with geometric and functional inequalities on manifolds, notably, the logarithmic Sobolev and energy-entropy inequalities [25], [26], [38]–[42]. Recall that a Riemannian manifold (M, g) is said to be a gradient Ricci soliton if there exists a smooth function ϕ on M and a constant λ R such that (cf. [7], [43])

(1.4) R ic ϕ ( g ) = R ic ( g ) + ϕ = λ g .

The notion is a generalisation of an Einstein manifold and has a fundamental role in the analysis of singularities of the Ricci flow [9], [36], [44], [45]. Taking trace from both sides of (1.4) and using the contracted Bianchi identity leads one to a simple form of (1.3) with constant coefficients.

Another prominent class of nonlinear equations rooted in conformal geometry and studied extensively in this setting are Yamabe-type equations (see, [4], [11], [46], [47]). In the context of smooth metric measure spaces these equations can be broadly formulated as (see, e.g., [35], [48], [49])

(1.5) w t = Δ ϕ w + A ( x , t ) w p + B ( x , t ) w .

Incidentally, the case A 1 , B 1 , p = 3 [ G ( w ) = w w 3 ] is the Allen–Cahn equation and the case A c , B c , p = 2 [ G ( w ) = c w ( 1 w ) with c > 0] is the Fisher-KKP equation (cf. [50]–[52]). Both these equations have been studied extensively in recent years due to the significance of the phenomenon they model and their huge applications in physics and other sciences. (For various geometric estimates and their consequences see [53], [54] and the references therein). A far reaching generalisation of (1.5) with a superposition of power-like nonlinearities consist of equations in the form

(1.6) w t = Δ ϕ w + j = 1 d A j ( x , t ) w p j + j = 1 d B j ( x , t ) w q j .

Here A j , B j (with 1 ≤ jd) are sufficiently smooth space-time dependent coefficients and p j ≥ 0, q j ≤ 0 real exponents (see [27], [28]). Other classes of equations generalising the above and close to (1.3) and (1.5) appear in the form (see, e.g., [27], [28], [34], [55], [56])

(1.7) w t = Δ ϕ w + A ( x , t ) Γ ( log w ) w p + B ( x , t ) w q + C ( x , t ) w ,

with p, q real exponents, A , B , C sufficiently smooth space-time dependent coefficients and Γ C 1 ( R , R ) . Some cases of particular interest for Γ = Γ(s) include power function, for example, s α (integer α ≥ 1), |s| α or |s| α−1 s (real α > 1) with different sign-changing, growth and singular behaviour as s → ±∞, or a superposition of such nonlinearities [28].

Another case of recent interest arises from iterated logarithms (cf. [55]) associated with a string of parameters d, k 1 , , k d N and β 1 , , β d R , specifically,

(1.8) Γ k 1 , , k d β 1 , , β d ( log w ) = | log k 1 w | β 1 | log k 2 w | β 2 | log k d w | β d ,

where log k w = log log k−1 w for k ≥ 2 and log1 w = logw. This can be considered but with due care, e.g., subject to the assumption of w being sufficiently large, specifically, with respect to iterated exponentials of k 1, …, k d (as otherwise the repeated logarithm is meaningless due to the possibility of log k−1 w being non-positive hence making log k w undefined). Naturally one can also consider variations of the same theme, for example, by replacing log k with either of its close relatives

(1.9) log k w = log log k 1 w  for  k 2 , log 1 w = | log w | ,

(1.10) log k + w = log + log k 1 + w  for  k 2 , log 1 + w = 1 + [ log w ] + .

However, one needs to observe that the function Γ thus obtained is only C 1 outside a discrete set (the zero sets of the functions log k−1 w) and hence does not lie in C 1 ( R , R ) as required.

Another related and yet more general form of Yamabe-type equations is the Einstein-scalar field Lichnerowicz equation (see Choquet-Bruhat [57], Chow [9] and Zhang [36]). In the context of smooth metric measure spaces a generalisation of the Einstein-scalar field Lichnerowicz equation with space-time dependent coefficients can be described as:

(1.11) w t = Δ ϕ w + A ( x , t ) w p + B ( x , t ) w q + C ( x , t ) w log w ,

(1.12) w t = Δ ϕ w + A ( x , t ) e 2 w + B ( x , t ) e 2 w + C ( x , t ) .

For gradient estimates, Harnack inequalities, Liouville type theorems and other related results in this direction see [27], [28], [30], [35], [49], [58], [59] and the references therein.

1.1 Plan and layout of paper

Let us conclude this introduction by briefly describing the plan of the paper. In Section 2 we present some background material on smooth metric measure spaces and the associated generalised Ricci curvature tensors as required for the development of the paper. In Section 3 we present the main results of the paper along with some related discussion. The remainder of the paper is then devoted to the detailed proofs of these results. In Section 4 we present the proof of the local Souplet-Zhang type gradient estimate in Theorem 3.1 and in Section 5 we give the proof of the local and global elliptic Harnack inequality in Theorem 3.3. In Section 6 we present the proof of the local Hamilton-type gradient estimate in Theorem 3.4 and in Section 7 we give the proof of the various Liouville-type results formulated in Theorems 3.7 and 3.8. In Section 8 we present the proof of the local Li-Yau differential Harnack estimate in Theorem 3.12 and in Section 9 we establish the local and global parabolic Harnack inequalities as formulated in Theorem 3.14. Section 10 is devoted to the proof of the general Liouville result in Theorem 3.16 and its subsequent corollaries.

For the convenience of the reader and future reference we describe at the end of this section some of the key notation and quantities used in the paper.

1.2 Notation

We denote by d = d(x, y) the geodesic distance between x, y in M. For the sake of the local estimates below we typically fix a reference point x 0 in M and denote by r = r(x) the geodesic radial variable measuring the distance between x, x 0 in M. We write B R ( x 0 ) for the closed geodesic ball in M centred at x 0 with radius R > 0 and we write Q R,T (x 0) for the compact parabolic space-time cylinder with lower base B R ( x 0 ) × { t 0 T } for t 0 R and height T > 0, specifically,

(1.13) Q R , T ( x 0 ) = B R ( x 0 ) × [ t 0 T , t 0 ] M × ( , ) .

When t 0 = T > 0 it is more convenient to write (1.13) as

(1.14) H R , T ( x 0 ) = B R ( x 0 ) × [ 0 , T ] M × [ 0 , T ] .

When the choices of x 0 and t 0 are clear from the context, in the interest of brevity we simply write B R , Q R,T and H R,T . We write X + = max(X, 0) and X = max(−X, 0).

2 Preliminaries and background

A smooth metric measure space or a weighted manifolds is a triple (M, g, dμ) where (M, g) is a smooth Riemannian manifold of dimension n ≥ 2 and dμ = eϕ dv g is the weighted measure associated with the potential ϕ (that is, the positive weight w = eϕ ) and the Riemannian volume measure dv g on M. The notion is an important one in geometric analysis and arises in various contexts (see, e.g., [12], [16]–[19], [22], [60], [61]).

Associated with the triple (M, g, dμ) there is a ϕ-Laplacian (also called the weighted, drifting or Witten Laplacian) defined by (1.2). The ϕ-Laplacian is a symmetric diffusion operator with respect to the invariant weighted measure dμ and reduces to the ordinary Laplacian (the Laplace-Beltrami operator) precisely when the potential ϕ is a constant function. By an application of the integration by parts formula it can be seen that for any compactly supported smooth functions u , w C 0 ( M ) we have

(2.1) M e ϕ w Δ ϕ u d v g = M e ϕ u , w d v g = M e ϕ u Δ ϕ w d v g .

As for the geometry and curvature properties of the triple (M, g, dμ) we introduce the generalised Ricci curvature tensor field on M by writing

(2.2) R ic ϕ m ( g ) = R ic ( g ) + ϕ ϕ ϕ m n ,

(the Bakry-Èmery m-Ricci curvature) where R ic ( g ) is the standard Riemannain Ricci curvature of g, ∇∇ϕ = Hess(ϕ) denotes the Hessian of ϕ, and mn is a constant (see, e.g., [38], [39], [62], [41], [43], [63], [64]). For the sake of clarity we point out that when m = n, by convention, ϕ is only allowed to be a constant, thus giving R ic ϕ m ( g ) = R ic ( g ) , whereas, we also allow for m = ∞ in which case by formally passing to the limit in (2.2) we get,

(2.3) R ic ϕ ( g ) = R ic ( g ) + ϕ R ic ϕ ( g ) .

The weighted Bochner-Weitzenböck formula in this context asserts that for every function w C 3 ( M ) we have the pointwise differential identity

(2.4) 1 2 Δ ϕ | w | 2 = | w | 2 + w , Δ ϕ w + R ic ϕ ( w , w ) .

Making note of the basic inequality (Δw)2/n ≤ |∇∇w|2 and by recalling the defining relation Δ ϕ w = Δw − ⟨∇ϕ, ∇w⟩ it is then easily seen that

(2.5) | w | 2 + ϕ ϕ m n ( w , w ) = | w | 2 + ϕ , w 2 m n ( Δ w ) 2 n + ϕ , w 2 m n ( Δ w ϕ , w ) 2 m = ( Δ ϕ w ) 2 m .

Subsequently it follows from (2.2), (2.4) and (2.5) that here we have the inequality

(2.6) 1 2 Δ ϕ | w | 2 1 m ( Δ ϕ w ) 2 + w , Δ ϕ w + R ic ϕ m ( w , w ) .

Now a curvature lower bound in the form R ic ϕ m ( g ) k g (with k R ) implies that the diffusion operator L = Δ ϕ satisfies the curvature-dimension condition CD(k, m) (see [38], [39], [41]–[43]). To elaborate on this last point further, we first recall the definition of the carre du champ operator Γ[L] associated with a Markov diffusion operator L. This is given explicitly by the formulation

(2.7) Γ [ L ] ( u , v ) = Γ 1 [ L ] ( u , v ) 1 2 [ L ( u v ) u L v v L u ] .

Higher order iterates are then defined inductively by replacing the product operation with Γ respectively (cf. [39]). As a matter of fact, the second order iterated carre du champ operator Γ2[L], can be seen to be given by,

(2.8) Γ 2 [ L ] ( u , v ) 1 2 [ L Γ ( u , v ) Γ ( u , L v ) Γ ( L u , v ) ] .

By a direct calculation, it is now observed that for the diffusion operator L = Δ ϕ as in (1.2) and with Γ[L] and Γ2[L] as in (2.7) and (2.8), we have the first and second order relations

(2.9) Γ [ L ] ( w , w ) = | w | 2 ,

and

(2.10) Γ 2 [ L ] ( w , w ) = 1 2 L | w | 2 w , L w ,

respectively. Hence, in light of (2.6), (2.9) and (2.10) it follows from the curvature lower bound R ic ϕ m ( g ) k g that here the iterated carre du champ operator Γ2[L] satisfies the inequality

(2.11) Γ 2 [ L ] ( w , w ) = 1 2 L | w | 2 w , L w 1 m ( L w ) 2 + R ic ϕ m ( w , w ) 1 m ( L w ) 2 + k | w | 2 = 1 m ( L w ) 2 + k Γ [ L ] ( w , w ) .

In contrast, a curvature lower bound in the form R ic ϕ ( g ) k g implies the curvature-dimension condition CD(k, ∞), in the sense that by virtue of (2.4), (2.9) and (2.10) one only has

(2.12) Γ 2 [ L ] ( w , w ) = 1 2 L | w | 2 w , L w = | w | 2 + R ic ϕ ( w , w ) k | w | 2 = k Γ [ L ] ( w , w ) .

3 Statement of the main results

Here we present the main results of the paper leaving the proofs, technicalities and further details involved to the subsequent sections. For the convenience of the reader and clarity of presentation, below we have grouped the results into four subsections.

3.1 Two gradient estimates of elliptic types for ( P )

We begin by presenting two estimates of elliptic type for positive solutions to ( P ) . These are estimates of Souplet-Zhang and Hamilton types respectively and the task is to exploit the role of the nonlinearity and the geometry of the triple (M, g, dμ). Here we fix a reference point x 0M and t 0 R , R > 2, T > 0. Since the solution w = w(x, t) > 0 and the parabolic space-time cylinder Q R,T (x 0) ⊂ M × (−∞, ∞) is compact, w is bounded away from zero and bounded from above in Q R,T (x 0) and so the quantities involving the nonlinearity and its derivatives in (3.1) are finite.

Theorem 3.1.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ ( g ) ( n 1 ) k g in B R for some k ≥ 0. Let w be a positive and bounded solution to ( P ) with 0 < wD. Then for all (x, t) in Q R/2,T with t > t 0T we have:

(3.1) | w | w ( x , t ) C 1 log w D ( x , t ) × k + 1 t t 0 + T + 1 R + [ γ Δ ϕ ] + R + sup Q R , T | G x ( t , x , w ) | w [ 1 log ( w / D ) ] 2 1 3 + sup Q R , T w [ 1 log ( w / D ) ] G w ( t , x , w ) + log ( w / D ) G ( t , x , w ) w [ 1 log ( w / D ) ] 2 + 1 2 .

Here C > 0 is a constant depending only on the dimension n, [ γ Δ ϕ ] + = max ( γ Δ ϕ , 0 ) where the functional quantity γ Δ ϕ is defined by

(3.2) γ Δ ϕ = max B 1 Δ ϕ r ( x ) ,

with B 1 = { x : d ( x , x 0 ) = 1 } .

The local estimate above has a global counterpart subject to the prescribed bounds in the theorem being global. The proof follows by passing to the limit R → ∞ in (3.1) and making note of the constants being independent of R whilst 1/R, [ γ Δ ϕ ] + / R 0 in the limit. The precise formulation of the estimate is given below.

Theorem 3.2.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ ( g ) ( n 1 ) k g in M for some k ≥ 0. Let w be a positive and bounded solution to ( P ) with 0 < wD. Then for all xM and t 0T < tt 0 we have:

(3.3) | w | w ( x , t ) C 1 log w D ( x , t ) × k + 1 t t 0 + T + sup M × [ t 0 T , t 0 ] | G x ( t , x , w ) | w [ 1 log ( w / D ) ] 2 1 3 + sup M × [ t 0 T , t 0 ] w [ 1 log ( w / D ) ] G w ( t , x , w ) + log ( w / D ) G ( t , x , w ) w [ 1 log ( w / D ) ] 2 + 1 2 .

Here C > 0 is a constant depending only on the dimension n.

One of the useful consequences of the estimates above is the following elliptic Harnack inequality for bounded positive solutions to ( P ) . Other implication will be presented subsequnetly.

Theorem 3.3.

Under the assumptions and local bounds in Theorem 3.1, for every x 1, x 2 in B R / 2 and t 0T < tt 0 with d = d(x 1, x 2), we have the Harnack inequality:

(3.4) w ( x 1 , t ) e D w ( x 2 , t ) e D α ,

where the exponent α = α(R) ∈ (0, 1) is given explicitly by

(3.5) exp d C k + 1 t t 0 + T + 1 R + [ γ Δ ϕ ] + R + sup Q R , T | G x ( t , x , w ) | w [ 1 log ( w / D ) ] 2 1 3 + sup Q R , T w [ 1 log ( w / D ) ] G w ( t , x , w ) + log ( w / D ) G ( t , x , w ) w [ 1 log ( w / D ) ] 2 + 1 2 .

If the prescribed bounds are global as in Theorem 3.2, then for every x 1, x 2 in M and t 0T < tt 0, the same inequality (3.4) holds now with the exponent α = α(M) given by

(3.6) exp d C k + 1 t t 0 + T + sup M × [ t 0 T , t 0 ] | G x ( t , x , w ) | w [ 1 log ( w / D ) ] 2 1 3 + sup M × [ t 0 T , t 0 ] w [ 1 log ( w / D ) ] G w ( t , x , w ) + log ( w / D ) G ( t , x , w ) w [ 1 log ( w / D ) ] 2 + 1 2 .

The second type of gradient estimates to be considered in this paper are of Hamilton type. Here again x 0M, t 0 R , R > 2, T > 0 are fixed. Unlike the Souplet-Zhang estimate in Theorem 3.1 here there is a choice of a free parameters α > 1 + β, β ≥ 0 in the estimate.

Theorem 3.4.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ ( g ) ( n 1 ) k g in B R for some k ≥ 0. Let w be a positive solution to ( P ) . Then for every β ≥ 0 and α > 1 + β > 0 and for all (x, t) in Q R/2,T with t > t 0T we have:

(3.7) | w | w 1 ( β + 2 ) / ( 2 α ) ( x , t ) C sup Q R , T w ( x , t ) ( β + 2 ) / ( 2 α ) × k + 1 t t 0 + T + 1 R + [ γ Δ ϕ ] + R + sup Q R , T | G x ( t , x , w ) | w 1 3 + sup Q R , T 2 w G w ( t , x , w ) [ 2 ( β + 2 ) / α ] G ( t , x , w ) w + 1 2 ,

where C = C(α, n) > 0 and γ Δ ϕ is as in (3.2).

Remark 3.1.

It is useful to point out that under the stronger curvature condition R ic ϕ m ( g ) ( m 1 ) k g one can remove in both the local estimates (3.1) and (3.7) the functional term

(3.8) ( [ γ Δ ϕ ] + / R ) 1 / 2 ,

that as will be seen in the proofs, arises from an application of the Wei-Wiley weighted Laplace comparison theorem under the lower bound on R ic ϕ ( g ) . (Under the stronger curvature condition, the constant C > 0 in (3.1) and (3.7) will depend on m.) Also note that for m = n where R ic ϕ n ( g ) = R ic ( g ) and Δ ϕ = Δ (here ϕ must be constant) our estimates and results on ( P ) are new even for the classical Laplace-Beltrami operator (see also [27], [28], [31]).

Again, the local estimate above has a global counterpart, when the asserted bounds in the theorem are global. The proof follows by passing to the limit R → ∞ in (3.7). This is the content of the following theorem.

Theorem 3.5.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ ( g ) ( n 1 ) k g in M for some k ≥ 0. Let w be a positive solution to ( P ) . Then for every β ≥ 0 and 1 + β < α and all xM and t 0T < tt 0 we have:

(3.9) | w | w 1 ( β + 2 ) / ( 2 α ) ( x , t ) C sup M × [ t 0 T , t 0 ] w ( x , t ) ( β + 2 ) / ( 2 α ) × k + 1 t t 0 + T + sup M × [ t 0 T , t 0 ] | G x ( t , x , w ) | w 1 3 + sup M × [ t 0 T , t 0 ] 2 w G w ( t , x , w ) [ 2 ( β + 2 ) / α ] G ( t , x , w ) w + 1 2 ,

where C = C(α, n) > 0.

3.2 Applications to Liouville-type results (I): R i c ϕ ( g ) 0

Let us now move on to discussing some applications of the gradient estimates above. Here we focus mainly on applications to Liouville and global constancy type results for both the parabolic and elliptic (stationary states) equations associated with ( P ) . Furthermore we assume G = G ( w ) .

Towards this end let us first suppose that w is a smooth bounded positive solution to the elliptic equation

(3.10) Δ ϕ w + G ( w ) = Δ w ϕ , w + G ( w ) = 0 .

Our main goal is to present two independent elliptic Liouville-type results for positive solutions to (3.10). Note that since a constant solution to (3.10) must be a zero of G , if G has no zeros w > 0 then a Liouville theorem can be seen as a non-existence result.

Theorem 3.6.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ ( g ) 0 in M. Let w be a positive and bounded solution to (3.10) with 0 < wD and assume [ 1 log ( w / D ) ] w G ( w ) + log ( w / D ) G ( w ) 0 . Then w is a constant. In particular G ( w ) = 0 .

The second Liouville-type result of interest as mentioned above is formulated below. The proof of both these theorems will be presented in Section 7.

Theorem 3.7.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ ( g ) 0 in M. Let w be a positive and bounded solution to (3.10) and assume for some β ≥ 0 and 1 + β < α we have [ 1 ( β / 2 + 1 ) / α ] G ( w ) w G ( w ) 0 . Then w is a constant. In particular G ( w ) = 0 .

Let us now discuss implications and some examples of the above theorems to specific nonlinearities of interest arising in various contexts within geometry and mathematical physics. For further discussion on this see the introduction to the paper.

Here we focus on nonlinearities in the “split” form G ( w ) = X ( w ) + w r Y ( log w ) where X = X ( w ) C 1 ( 0 , ) , r is a real exponent and Y = Y ( s ) = Y ( log w ) C 1 ( , ) . The motivation behind considering such types on nonlinearities was discussed earlier in Section 1 (see also Theorem 3.10 below). The next result is an application of Theorem 3.6.

Theorem 3.8.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ ( g ) 0 in M. Let w be a positive bounded solution to the nonlinear elliptic equation

(3.11) Δ ϕ w + X ( w ) + w r Y ( log w ) = 0 .

Suppose X′(w) ≤ 0 and X(w) − wX′(w) ≥ 0. Moreover, when Y ≢ 0, depending on w satisfying w ≥ 1 or 0 < w ≤ 1 everywhere on M, assume that:

(3.12) ( Y 1 ) r 1 ,  w 1 , Y ( s ) 0  and Y ( s ) 0  for all  s 0 ,

(3.13) o r , ( Y 2 ) r min ( γ , 1 ) ,  0 < w 1 , Y ( s ) 0 ,  Y ( s ) 0  and s Y ( s ) γ Y ( s )  for all  s 0  and some  γ 0 .

Then w is a constant. In particular X(w) + w r Y(log w) = 0.

Regarding the assumptions in the theorem note that the examples X ( w ) = A w p (with A 0 and p ≤ 0) and X ( w ) = B w q (with B 0 and q ≥ 1) satisfy the assumptions on X. Moreover, the example Y(s) = |s| γ (γ > 1) satisfies Y 2 , Y(s) = −|s| γ satisfies Y 1 and the example Y(s) = −s γ (γ ≥ 1 odd integer) satisfies both Y 1 and Y 2 . (Compare these to Theorem 4.3 in [28] and the choice X = X(w) in (3.14) below.)

As another example consider the case where the nonlinearity is a superposition of power-like terms in the form

(3.14) X ( w ) = j = 1 N A j w p j + j = 1 N B j w q j .

Here A j , B j are constant coefficients and p j , q j for 1 ≤ jN are real exponents. Then we can formulate the following result which improves and extends earlier results on Yamabe-type problems (cf. [11], [27], [28], [35], [49]). This is an application of Theorem 3.7.

Theorem 3.9.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ ( g ) 0 in M. Let w be a positive bounded solution to the nonlinear elliptic equation

(3.15) Δ ϕ w + j = 1 N A j w p j + j = 1 N B j w q j = 0 .

Assume A j 0 , B j 0 and p j ≤ 1 − (β/2 + 1)/α, q j ≥ 1 − (β/2 + 1)/α for 1 ≤ jN and for some β ≥ 0 and 1 + β < α. Then w must be a constant and G ( w ) = 0 .

As a final application consider G ( w ) = A w p + B w q + w r Y ( log w ) with real exponents p, q, r, constant coefficients A , B and Y C 1 ( R ) . Then Theorem 3.7 leads to the following.

Theorem 3.10.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ ( g ) 0 in M. Let w be a positive bounded solution to the nonlinear elliptic equation

(3.16) Δ ϕ w + A w p + B w q + w r Y ( log w ) = 0 .

Assume that along the solution w we have the inequality Y′ + [(β/2 + 1)/α + r − 1]Y ≤ 0 along with A 0 , B 0 , p ≤ 1 − (β/2 + 1)/α and q ≥ 1 − (β/2 + 1)/α for some β ≥ 0 and 1 + β < α. Then w must be a constant and A w p + B w q + w r Y ( log w ) = 0 .

Note that the assumption Y′ + [(β/2 + 1)/α + r − 1]Y ≤ 0 in Theorem 3.10 is implied by either of ( H 1 ) or ( H 2 ) below on Y = Y ( s ) = Y ( log w ) C 1 ( , ) :

(3.17) ( H 1 ) r 1 ( β / 2 + 1 ) / α ,  w 1 , Y 0  and  Y 0  for all  s 0 ,

(3.18) o r , ( H 2 ) r 1 ( β / 2 + 1 ) / α ,  0 < w 1 , Y 0  and  Y 0  for all  s 0 .

We end this section with an application of the estimates above to parabolic Liouville-type theorems. Here by an ancient solution w = w(x, t) to ( P ) we mean a solution defined on M for all negative times, that is, for all (x, t) with xM and −∞ < t < 0.

Theorem 3.11.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ ( g ) 0 in M. Let G ( w ) 0 , G ( w ) w G ( w ) 0 and G ( w ) a for some a > 0 and all w > 0. Then the equation

(3.19) w t Δ ϕ w = G ( w ) ,

does not have any positive ancient solutions satisfying w ( x , t ) = e o ( r ( x ) + | t | ) .

3.3 A parabolic Li-Yau type gradient estimate for ( P )

Let us turn to our final gradient estimate for positive solutions to ( P ) . Towards this end it is convenient to introduce further notation that will appear in different stages of the analysis and serve as bounds in the estimates. In order to describe these, for given G as above of class C 2 and constant α we set,

(3.20) A G α ( t , x , w ) = α w G w w ( t , x , w ) + G w ( t , x , w ) w 1 G ( t , x , w ) + ,

(3.21) B G α ( t , x , w ) = | α G x w ( t , x , w ) w 1 G x ( t , x , w ) | ,

(3.22) C G ( t , x , w ) = G w ( t , x , w ) w 1 G ( t , x , w ) + ,

(3.23) D G ( t , x , w ) = w 1 Δ ϕ G x ( t , x , w ) + .

Here as before subscripts denote partial derivatives. Moreover G x : x G ( t , x , w ) denotes the function obtained by freezing the variables t, w and viewing G as a function of x only. (Thus in particular we speak of G x and Δ ϕ G x .) Having the above notation in place we now define the four pairs of γ-quantities associated with a given w = w(x, t) (xM, 0 ≤ tT) by writing for fixed x 0M, R > 0 and T > 0:

(3.24) γ A G , α ( R ) = sup H R , T A G α ( t , x , w ) , γ A G , α = sup M × [ 0 , T ] A G α ( t , x , w ) ,

(3.25) γ B G , α ( R ) = sup H R , T B G α ( t , x , w ) , γ B G , α = sup M × [ 0 , T ] B G α ( t , x , w ) ,

(3.26) γ C G ( R ) = sup H R , T C G ( t , x , w ) , γ C G = sup M × [ 0 , T ] C G ( t , x , w ) ,

(3.27) γ D G ( R ) = sup H R , T D G ( t , x , w ) , γ D G = sup M × [ 0 , T ] D G ( t , x , w ) .

For the first estimate we fix x 0M, R > 0, T > 0 and formulate an upper bound for the Harnack quantity [see the left-hand side of (3.28)] on the cylinder B R × ( 0 , T ] .

Theorem 3.12.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ m ( g ) ( m 1 ) k g in B 2 R for some mn and k ≥ 0. Let w be a smooth positive solution to ( P ) . Then for every α > 1, ɛ ∈ (0, 1) and all (x, t) ∈ H R,T with t > 0 we have

(3.28) | w | 2 α w 2 t w w + G w ( x , t ) m α 2 1 R 2 m c 1 2 α 2 4 ( α 1 ) + c 2 + ( m 1 ) c 1 ( 1 + R k ) + 2 c 1 2 + 1 t + γ C G ( 2 R ) + m α 2 m α ( m 1 ) k + γ A G , α ( 2 R ) / 2 2 2 ( 1 ε ) ( α 1 ) 2 + 3 3 m γ B G , α ( 2 R ) 4 2 5 ε α ( α 1 ) 2 1 / 3 + γ D G ( 2 R ) 1 / 2 .

The constants c 1, c 2 > 0 in (3.28) are those appearing in the bounds (8.15) in Lemma 8.3 and the γ-quantities are as in (3.24)(3.27) (with 2R replacing R).

The local estimate above has a global in space counterpart subject to the prescribed bounds in the theorem being global in space. The proof follows by passing R → ∞ in (3.28) and taking into account the vanishing of the term

(3.29) 1 R 2 m c 1 2 α 2 4 ( α 1 ) + c 2 + ( m 1 ) c 1 ( 1 + R k ) + 2 c 1 2

in the limit due to the constants being independent of R. The precise formulation of this global estimate is given in the theorem below.

Theorem 3.13.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ m ( g ) ( m 1 ) k g in M for some mn and k ≥ 0. Let w be a smooth positive solution to ( P ) . Then for every α > 1, ɛ ∈ (0, 1) and all xM, 0 < tT we have

(3.30) | w | 2 α w 2 t w w + G w ( x , t ) m α 2 t + m α 2 m α ( m 1 ) k + γ A G , α / 2 2 2 ( 1 ε ) ( α 1 ) 2 + 3 3 m γ B G , α 4 2 5 ε α ( α 1 ) 2 1 / 3 + γ D G 1 / 2 + m α 2 γ C G .

Here the γ-quantities on the right in (3.30) are as given by (3.24)(3.27).

An immediate consequence of the local and global estimates in Theorems 3.12 and 3.13 is the following parabolic Harnack inequality on the solutions.

Theorem 3.14.

Under the assumptions of Theorem 3.12 let w be a positive solution to ( P ) . Then for every (x 1, t 1), (x 2, t 2) in H R,T with t 2 > t 1 > 0 and α > 1,

(3.31) w ( x 2 , t 2 ) w ( x 1 , t 1 ) exp [ ( t 2 t 1 ) H α L ( x 1 , x 2 , t 2 t 1 ) ] t 2 t 1 m α / 2 .

Here H is a constant depending only on the bounds given in Theorem 3.12 (see (9.1) and (9.2)) and L is given by

(3.32) L ( x 1 , x 2 , t 2 t 1 ) = inf ζ Γ 1 4 ( t 2 t 1 ) 0 1 | ζ ̇ ( t ) | 2 d t ,

where Γ = Γ ( x 1 , x 2 ) = ζ C 1 ( [ 0 , 1 ] , M ) : ζ ( [ 0 , 1 ] ) B R , ζ ( 0 ) = x 1 , ζ ( 1 ) = x 2 . If the bounds are global as in Theorem 3.13 then the Harnack inequality (3.31) is global too.

3.4 Applications to Liouville-type results (II): R i c ϕ m ( g ) 0

We complement the Liouville-type results given earlier by invoking the Li-Yau gradient estimates above which were obtained under a different curvature condition. We begin by the following theorem on the stationary version of ( P ) that is of independent interest.

Theorem 3.15.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ m ( g ) ( m 1 ) k g in M. Let w be a smooth positive solution to (3.10). Then, for every α > 1 and ɛ ∈ (0, 1) the following global estimate holds on M:

(3.33) | w | 2 α w 2 + G ( w ) w m α 2 ( m 1 ) k + γ A G , α / 2 ( α 1 ) 1 ε + γ C G .

Theorem 3.16.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ m ( g ) 0 in M. Let w be a smooth positive solution to (3.10) and assume that along the solution w we have G ( w ) 0 , G ( w ) w G w ( w ) 0 and

(3.34) G ( w ) w G w ( w ) + α w 2 G w w ( w ) 0 ,

for some α > 1, everywhere on M. Then w must be a constant and G ( w ) = 0 .

Theorem 3.16 leads to the following conclusion on Yamabe equations. As indicated earlier, since a constant solution to (3.10) must be a zero of G , when G has no zeros w > 0, a Liouville theorem can be seen as a non-existence result for positive solutions. In the following theorem this happens precisely when A j > 0 for at least one 1 ≤ jd. Compare also with Theorem 3.9 and the choices of curvature conditions there and here.

Theorem 3.17.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ m ( g ) 0 in M. Let w be a positive smooth solution to the equation

(3.35) Δ ϕ w + j = 1 N A j w p j = 0 .

If A j 0 and p j ≤ 1 for all 1 ≤ jN then w must be a constant and A j w p j = 0 .

As another application consider a nonlinearity in the form of a superposition of a logarithmic and a power-like nonlinearity with constant coefficients A , B , exponent s and Y C 2 ( R ) in the form

(3.36) G ( w ) = A w Y ( log w ) + B w s .

The following theorem now directly results from Theorem 3.16.

Theorem 3.18.

Let (M, g, dμ) with dμ = eϕ dv g be a complete smooth metric measure space satisfying R ic ϕ m ( g ) 0 . Let w be a positive smooth solution to the equation

(3.37) Δ ϕ w + A w Y ( log w ) + B w s = 0 .

Assume that along the solution w we have Y ≥ 0, Y′ ≤ 0 and αY″ + (α − 1)Y′ ≥ 0 for some α > 1. Furthermore assume A , B 0 and s ≤ 1. Then w must be a constant and A w Y ( log w ) + B w s = 0 .

4 Proof of the local Souplet-Zhang estimate in Theorem 3.1

We now attend to the proof of the results formulated in Section 3. Throughout this section we confine to the proof of Theorem 3.1 and all of its necessary ingredients. The first two subsections develop the necessary apparatus, including an evolution inequality on a quantity (below called H) built out of the solution w, and the last section completes the proof by putting these together, using a localisation argument and finally maximum principle.

4.1 Evolution inequality for H = |∇[log(1 − h)]|2 with h = log(w/D)

From the positive solution w satisfying the bounds 0 < wD we first define a non-positive function by putting h = log(w/D) and then H = |∇h|2/(1 − h)2. The evolution of H under the equation ( P ) is the subject of the next lemma.

Lemma 4.1.

Let w be a positive and bounded solution to ( P ) with 0 < wD. Put h = log(w/D) and let H = |∇h|2/(1 − h)2. Then H satisfies the evolution equation

(4.1) [ Δ ϕ t ] H = 2 R ic ϕ ( g ) ( h , h ) ( 1 h ) 2 + 2 h h , H 1 h + 2 ( 1 h ) H 2 + 2 2 h 1 h + h h ( 1 h ) 2 2 2 h , G x ( t , x , D e h ) D e h ( 1 h ) 2 2 H G w ( t , x , D e h ) + h G ( t , x , D e h ) D e h ( 1 h ) .

In particular, if R ic ϕ ( g ) ( n 1 ) k g for some k ≥ 0, then

(4.2) [ Δ ϕ t ] H 2 ( n 1 ) k H + 2 h h , H 1 h + 2 ( 1 h ) H 2 2 h , G x ( t , x , D e h ) D e h ( 1 h ) 2 2 H G w ( t , x , D e h ) + h G ( t , x , D e h ) D e h ( 1 h ) .

Proof.

As w is a solution to ( P ) and 0 < wD it is easily seen that the non-positive function h = log(w/D) satisfies the equation

(4.3) t h = Δ ϕ h + | h | 2 + D 1 e h G ( t , x , D e h ) .

As a result, by noting of G ( t , x , D e h ) = G x ( t , x , D e h ) + D e h h G w ( t , x , D e h ) it follows that

(4.4) t | h | 2 = 2 h , t h = 2 h , Δ ϕ h + | h | 2 + D 1 e h G ( t , x , D e h ) = 2 h , Δ ϕ h + 2 h , | h | 2 + 2 h , G x ( t , x , D e h ) D e h + 2 | h | 2 G w ( t , x , D e h ) G ( t , x , D e h ) D e h .

Now moving on to the function H = |∇h|2/(1 − h)2 = |∇log(1 − h)|2 it is easily seen that t H = t | h | 2 / ( 1 h ) 2 + 2 | h | 2 t h / ( 1 h ) 3 and so making note of (4.3) and (4.4) we can write

(4.5) t H = 2 h , Δ ϕ h ( 1 h ) 2 + 2 h , | h | 2 ( 1 h ) 2 + 2 h , G x ( t , x , D e h ) D e f ( 1 h ) 2 + 2 | h | 2 ( 1 h ) 2 G w ( t , x , D e h ) G ( t , x , D e h ) D e h + 2 | h | 2 Δ ϕ h ( 1 h ) 3 + 2 | h | 4 ( 1 h ) 3 + 2 | h | 2 G ( t , x , D e h ) D e h ( 1 h ) 3 .

Likewise we have ∇H = [∇|∇h|2]/(1 − h)2 + [2|∇h|2h]/(1 − h)3 and so by recalling the formulation Δ ϕ H = ΔH − ⟨∇ϕ, ∇H⟩ it follows that

(4.6) Δ ϕ H = Δ ϕ | h | 2 ( 1 h ) 2 + 4 h , | h | 2 ( 1 h ) 3 + 2 | h | 2 Δ ϕ h ( 1 h ) 3 + 6 | h | 4 ( 1 h ) 4 .

Putting (4.5) and (4.6) together and taking into account the necessary cancellations give

(4.7) [ Δ ϕ t ] H = Δ ϕ | h | 2 ( 1 h ) 2 2 h , Δ ϕ h ( 1 h ) 2 2 h , | h | 2 ( 1 h ) 2 2 | h | 4 ( 1 h ) 3 + 4 h , | h | 2 ( 1 h ) 3 + 6 | h | 4 ( 1 h ) 4 2 h , G x ( t , x , D e h ) D e h ( 1 h ) 2 2 | h | 2 ( 1 h ) 2 G w ( t , x , D e h ) G ( t , x , D e h ) D e h 2 | h | 2 G ( t , x , D e h ) D e h ( 1 h ) 3 .

Now by making use of the weighted Bochner–Weitzenböck formula we obtain

(4.8) [ Δ ϕ t ] H = 2 | h | 2 ( 1 h ) 2 + 2 R ic ϕ ( g ) ( h , h ) ( 1 h ) 2 2 h , | h | 2 ( 1 h ) 2 + 4 h , | h | 2 ( 1 h ) 3 2 | h | 4 ( 1 h ) 3 + 6 | h | 4 ( 1 h ) 4 2 h , G x ( t , x , D e h ) D e h ( 1 h ) 2 2 | h | 2 ( 1 h ) 2 G w ( t , x , D e h ) G ( t , x , D e h ) D e h 2 | h | 2 G ( t , x , D e h ) D e h ( 1 h ) 3 ,

and therefore a rearrangement of terms and basic considerations leads to

(4.9) [ Δ ϕ t ] H = 2 R ic ϕ ( g ) ( h , h ) ( 1 h ) 2 + 2 h 1 h + h h ( 1 h ) 2 2 + 2 | h | 4 ( 1 h ) 3 2 h , | h | 2 ( 1 h ) 2 4 | h | 4 ( 1 h ) 3 + 2 h , | h | 2 ( 1 h ) 3 + 4 | h | 4 ( 1 h ) 4 2 h , G x ( t , x , D e h ) D e h ( 1 h ) 2 2 | h | 2 G ( t , x , D e h ) D e h ( 1 h ) 3 2 | h | 2 ( 1 h ) 2 G w ( t , x , D e h ) G ( t , x , D e h ) D e h .

Finally by making note of the relation (1 − h)3⟨∇h, ∇H⟩ = (1 − h)⟨∇h, ∇|∇h|2⟩ + 2|∇h|4 the expression on the second line on the right simplifies to 2h⟨∇h, ∇H⟩/(1 − h) and so a reference to H = |∇h|2/(1 − h)2 leads at once to the desired conclusion.

The inequality (4.2) now follows from the above by using the Ricci curvature lower bound R ic ϕ ( g ) ( n 1 ) k g and the non-negativity of the quadratic term on the first line on the right. The proof is thus complete. □

4.2 Construction of space-time cut-offs and cylindrical localisation

In order to prove the estimate in Theorem 3.1 we first need to establish a local version of the evolution inequality in Lemma 4.1 suitable for the application of maximum principle. This will be achieved through the use of suitable space-time cut-off functions. Towards this end, let us fix x 0M, t 0 R , R, T > 0 and then τ ∈ (t 0T, t 0]. The following standard lemma grants the existence of a smooth function η ̄ = η ̄ ( r , t ) of two real variables r ≥ 0 and t 0Ttt 0 respectively satisfying a set of useful bounds and properties for carrying out this cylindrical localisation procedure (see, e.g., [3], [23], [65]).

Lemma 4.2.

Fix t 0 R and let R, T > 0. Given τ ∈ (t 0T, t 0] there exists a smooth function η ̄ : [ 0 , ) × [ t 0 T , t 0 ] R such that the following properties hold:

  1. s u p p η ̄ ( r , t ) [ 0 , R ] × [ t 0 T , t 0 ] and 0 η ̄ ( r , t ) 1 in [0, R] × [t 0T, t 0],

  2. η ̄ = 1 in [0, R/2] × [τ, t 0] and η ̄ / r = 0 in [0, R/2] × [t 0T, t 0], respectively,

  3. there exists c > 0 such that

    (4.10) η ̄ t c η ̄ 1 / 2 τ t 0 + T ,

    in [0, ∞) × [t 0T, t 0] and η ̄ ( r , t 0 T ) = 0 for all r ∈ [0, ∞).

  4. c a η ̄ a / R η ̄ / r 0 and | 2 η ̄ / r 2 | c a η ̄ a / R 2 hold on [0, ∞) × [t 0T, t 0] for every 0 < a < 1 and some c a > 0.

Having the above lemma at our disposal we now move on to introducing a smooth space-time cut-off function η = η(x, t) by setting, for R ≥ 2, T > 0 and t 0T < τt 0 (the reason for the choice R ≥ 2 will be clear later),

(4.11) η ( x , t ) = η ̄ ( r ( x ) , t ) , ( x , t ) M × [ t 0 T , t 0 ] .

It is plain that η is supported in the compact space-time cylinder Q R,T M × [t 0T, t 0]. Furthermore in view of (4.11) we have Δ f η = η ̄ r r | r | 2 + η ̄ r Δ f r and t η = η ̄ t .

Let us now move on to the following useful product identity for the action of the weighted heat operator on the space-time localised function ηw.

Lemma 4.3.

Let η = η(x, t) be as above and let H = H(x, t) be a space-time function of class C 2 . Then we have

(4.12) [ Δ ϕ t ] ( η H ) = η [ Δ ϕ t ] H + 2 [ η , ( η H ) | η | 2 H ] / η + H [ Δ ϕ t ] η .

Proof.

Firstly [Δ ϕ − ∂ t ](ηH) = η ϕ − ∂ t ]H + 2⟨∇H, ∇η⟩ + H ϕ − ∂ t ]η as is easily seen by direct differentiation. Next, ⟨∇η, ∇(ηH)⟩ = ⟨∇η, Hη + ηH⟩ = H|∇η|2 + η⟨∇η, ∇H⟩ and therefore using the latter relation to substitute for the middle term 2⟨∇H, ∇η⟩ in the first identity gives the desired result. □

4.3 Finalising the proof of Theorem 3.1

Let us now fix τ ∈ (t 0T, t 0] and put η ( x , t ) = η ̄ ( r ( x ) , t ) with η ̄ as in Lemma 4.2. We show that the desired estimate holds at all points (x, τ) with d(x, x 0) ≤ R/2. The arbitrariness of τ will then give the assertion for all (x, t) in Q R/2,T with t > t 0T. Now starting from the inequality (4.2) in Lemma 4.1 we have

(4.13) [ Δ ϕ t ] H 2 ( 1 h ) H 2 2 ( n 1 ) k H + 2 h h , H 1 h 2 h , G x ( t , x , D e h ) D e h ( 1 h ) 2 2 H G w ( t , x , D e h ) + h G ( t , x , D e h ) D e h ( 1 h ) .

Hence for the localised function ηH we have upon invoking Lemma 4.3 the inequality

(4.14) [ Δ ϕ t ] ( η H ) 2 h h 1 h + 2 η η , ( η H ) H 2 h h 1 h + 2 η η , η + 2 ( 1 h ) η H 2 + H [ Δ ϕ t 2 ( n 1 ) k ] η 2 η h , G x ( t , x , D e h ) D e h ( 1 h ) 2 2 η H G w ( t , x , D e h ) + h G ( t , x , D e h ) D e h ( 1 h ) .

Assume that the localised function ηH is maximal at the point (x 1, t 1) in the compact space-time cylinder {(x, t): d(x, x 0) ≤ R, t 0Ttτ}. Suppose also that x 1 is not in the cut locus of M by Calabi’s argument [15] and that (ηH)(x 1, t 1) > 0 as otherwise the result is trivial with w(x, τ) ≤ 0 for all d(x, x 0) ≤ R/2. It then follows that t 1 > t 0T at (x 1, t 1) and subsequently by maximum principle

(4.15) t ( η H ) 0 , ( η H ) = 0 , Δ ( η H ) 0 .

In particular at (x 1, t 1) we have [Δ ϕ − ∂ t ](ηH) ≤ 0 and therefore from (4.14) we obtain the inequality

(4.16) 2 ( 1 h ) η H 2 H 2 h h 1 h + 2 η η , η H Δ ϕ η + H t η + 2 ( n 1 ) k H η + 2 η h , G x ( t , x , D e h ) D e h ( 1 h ) 2 + 2 η H G w ( t , x , D e h ) + h G ( t , x , D e h ) D e h ( 1 h ) .

As a result dividing through by 2(1 − h) ≥ 0 it follows that

(4.17) η H 2 H h h 1 h + η η , η 1 h + H [ Δ ϕ + t + 2 ( n 1 ) k ] η 2 ( 1 h ) + η h , G x ( t , x , D e h ) D e h ( 1 h ) 3 + η H G w ( t , x , D e h ) 1 h + h G ( t , x , D e h ) D e h ( 1 h ) 2 .

The goal is now to utilise (4.17) to establish the required estimate at (x, τ). Towards this end we proceed by considering two alternatives: d(x 1, x 0) ≤ 1 and d(x 1, x 0) ≥ 1.

Alternative 1.

(d(x 1, x 0) ≤ 1). Since here η is a constant function in the space [for all x with d(x, x 0) ≤ R/2 and R ≥ 2 by property (ii)] the terms involving space derivatives of η at (x 1, t 1) vanish (that is, ∇η = 0, Δ ϕ η = 0 and t η = η ̄ t ). Thus (4.17) reduces to

(4.18) η H 2 H [ | t η | + 2 ( n 1 ) k η ] 2 ( 1 h ) + η | h , G x ( t , x , D e h ) | D e h ( 1 h ) 3 + η H G w ( t , x , D e h ) 1 h + h G ( t , x , D e h ) D e h ( 1 h ) 2 + ,

and so

(4.19) η H 2 η H 2 ( 1 h ) | t η ̄ | η ̄ + ( n 1 ) k η H ( 1 h ) + η H | G x ( t , x , D e h ) | D e h ( 1 h ) 2 + η H G w ( t , x , D e h ) 1 h + h G ( t , x , D e h ) D e h ( 1 h ) 2 + .

Next by an application of Cauchy-Schwarz and Young’s inequality and after rearranging terms it follows that for suitable C = C(n) > 0 we have

(4.20) η H 2 C 1 ( τ t 0 + T ) 2 + k 2 + | G x ( t , x , D e h ) | D e h ( 1 h ) 2 4 / 3 + G w ( t , x , D e h ) 1 h + h G ( t , x , D e h ) D e h ( 1 h ) 2 + 2 .

As η ≡ 1, when d(x, x 0) ≤ R/2 and τtt 0 (and so in particular for when t = τ) by (i), we have H ( x , τ ) = η H ( x , τ ) η H ( x 1 , t 1 ) η H 2 4 ( x 1 , t 1 ) . Hence recalling H = |∇h|2/(1 − h)2 and the relation h = log(w/D), we arrive at the bound at (x, τ)

(4.21) h 1 h C 1 τ t 0 + T + k + sup Q R , T | G x ( t , x , D e h ) | D e h ( 1 h ) 2 1 / 3 + sup Q R , T D e h ( 1 h ) G w ( t , x , D e h ) + h G ( t , x , D e h ) D e h ( 1 h ) 2 + 1 / 2 .

This together with the arbitrariness of τ > t 0T is now seen to give a special case of (3.1).

Alternative 2.

(d(x 1, x 0) ≥ 1). Upon referring to the right-hand side of (4.17), and noting the properties of η ̄ as listed in Lemma 4.2 we proceed onto bounding each of the four terms on the right-hand side on (4.17).

  1. Towards this end dealing with the first term first, we have

    (4.22) H h h 1 h + η η , η 1 h H h | h | 1 h + | η | η | η | 1 h H h H + | η | η | η | 1 h H η η 1 / 4 H | h | 1 h | η | η 3 / 4 + | η | 2 η 3 / 2 η H 2 5 + C R 4 .

  2. Here we utilise the Wei-Wylie weighted Laplacian comparison theorem along with the lower curvature bound on R ic ϕ ( g ) . Indeed recalling r(x) = d(x, x 0), 1 ≤ d(x 1, x 0) ≤ R and R ic ϕ ( g ) ( n 1 ) k g with k ≥ 0, upon referring to (4.11) we can write

    (4.23) Δ ϕ η = η ̄ r r | r | 2 + η ̄ r Δ ϕ r ( η ̄ r r + η ̄ r [ γ Δ ϕ + ( n 1 ) k ( R 1 ) ] )

    where we have used Theorem 3.1 in Ref. [66] to write Δ ϕ r [ γ Δ ϕ ] + + ( n 1 ) k ( R 1 ) whenever 1 ≤ rR, t 0Ttt 0 [thus in particular at the space-time point (x 1, t 1)]. Note that here we have used (ii) [ η ̄ r = 0 when 0 ≤ rR/2] and (iv) [ η ̄ r 0 when 0 ≤ r < ∞] in Lemma 4.2. Hence putting the above together, by an application of Young’s inequality, we can write

    (4.24) H ( Δ ϕ η ) 1 h η H 1 h | η ̄ r r | η ̄ + | η ̄ r | η ̄ [ γ Δ ϕ ] + + ( n 1 ) k ( R 1 ) C η H 1 h 1 R 2 + [ γ Δ ϕ ] + R + k η H 2 10 + C 1 R 4 + [ γ Δ ϕ ] + 2 R 2 + k 2 ,

    and in a similar way using (iii) in Lemma 4.2, t η = η ̄ t c η / ( τ t 0 + T ) , and therefore [ 2 ( n 1 ) k + t ] η η [ 2 ( n 1 ) k + c / ( τ t 0 + T ) ] . An application of Young’s inequality now gives [ H / ( 1 h ) ] [ 2 ( n 1 ) k + t ] η η H 2 / 10 + C 1 / ( τ t 0 + T ) 2 + k 2 . Hence returning the second term on the right in (4.17) and combining the above we have

    (4.25) H [ Δ ϕ + t + 2 ( n 1 ) k ] η 2 ( 1 h ) η H 2 5 + C 1 R 4 + [ γ Δ ϕ ] + 2 R 2 + 1 ( τ t 0 + T ) 2 + k 2 .

  3. In much the same way regarding the third term involving G ( t , x , D e h ) we have

    (4.26) η h , G x ( t , x , D e h ) D e h ( 1 h ) 3 η | h | | G x ( t , x , D e h ) | D e h ( 1 h ) 3 = η H | G x ( t , x , D e h ) | D e h ( 1 h ) 2 η H 2 5 + C | G x ( t , x , D e h ) | D e h ( 1 h ) 2 4 / 3 .

  4. Likewise for the subsequent terms, upon noting −1 ≤ h/(1 − h) ≤ 0, h ≤ 0 and 0 ≤ η ≤ 1, we have

    (4.27) η H G w ( t , x , D e h ) 1 h + h G ( t , x , D e h ) D e h ( 1 h ) 2 = η H D e h ( 1 h ) G w ( t , x , D e h ) + h G ( t , x , D e h ) D e h ( 1 h ) 2 η H 2 5 + C D e h ( 1 h ) G w ( t , x , D e h ) + h G ( t , x , D e h ) D e h ( 1 h ) 2 + 2 .

Now referring to (4.17) noting the inequality 1 − h ≥ 1 and making use of the bounds obtained in (4.22)(4.27), it follows the following upper bound holds for ηH 2 at (x 1, t 1),

(4.28) η H 2 C k 2 + 1 ( τ t 0 + T ) 2 + 1 R 4 + [ γ Δ ϕ ] + 2 R 2 + sup Q R , T | G x ( t , x , D e h ) | D e h ( 1 h ) 2 4 / 3 + sup Q R , T D e h ( 1 h ) G w ( t , x , D e h ) + h G ( t , x , D e h ) D e h ( 1 h ) 2 + 2 .

Recalling the maximality of ηH at (x 1, t 1) along with η ≡ 1 when d(x, x 0) ≤ R/2 and τtt 0, it follows that H 2(x, τ) = (η 2 H 2)(x, τ) ≤ (η 2 H 2)(x 1, t 1) ≤ (ηH 2)(x 1, t 1) when d(x, x 0) ≤ R/2. hence upon noting H = |∇h|2/(1 − h)2, the above gives

(4.29) log w 1 log ( w / D ) C k + 1 τ t 0 + T + 1 R + [ γ Δ ϕ ] + R + sup Q R , T | G x ( t , x , D e h ) | D e h ( 1 h ) 2 1 / 3 + sup Q R , T D e h ( 1 h ) G w ( t , x , D e h ) + h G ( t , x , D e h ) D e h ( 1 h ) 2 1 / 2 .

Thus in either case we have shown the estimate is true at (x, τ). The desired conclusion now follows from the arbitrariness of τ ∈ (t 0T, t 0]. The proof is thus complete.

5 Proof of the elliptic Harnack inequality in Theorem 3.3

We now come to the proof of the Harnack inequality in Theorem 3.3. To this end pick x 1, x 2 in M and t 0T < t < t 0. Let ζ = ζ(s) with 0 ≤ s ≤ 1 be a shortest geodesic curve with respect to g, lying completely in the closed ball B R / 2 M and joining the points x 1, x 2, specifically, ζ(0) = x 1 and ζ(1) = x 2. Let us also put d = d(x 1, x 2).

Henceforth we shall assume the G -terms in the theorem are finite (else α(R) = 0 and the desired inequality is trivially true). Now utilising the estimate (3.1) in Theorem 3.1 we can write

(5.1) log 1 h ( x 2 , t ) 1 h ( x 1 , t ) = 0 1 d d s log [ 1 h ( ζ ( s ) , t ) ] d s = 0 1 h ( ζ ( s ) , t ) , ζ ( s ) 1 h ( ζ ( s ) , t ) d s

0 1 | h ζ | 1 h d s sup Q R / 2 , T | h | 1 h 0 1 | ζ | d s = d sup Q R / 2 , T | h | 1 h

d C k + 1 t t 0 + T + 1 R + [ γ Δ ϕ ] + R + sup Q R , T | G x ( t , x , D e h ) | D e h ( 1 h ) 2 1 / 3 + sup Q R , T D e h ( 1 h ) G w ( t , x , D e h ) + h G ( t , x , D e h ) D e h ( 1 h ) 2 + 1 / 2 ,

where in the language of (3.5) the last expression on the right is −log α(R). Moreover here we have used |∇h|/(1 − h) = |∇log w|/[1 − log(w/D)]. Therefore exponentiating (5.1) results in

(5.2) log [ e D / w ( x 2 , t ) ] log [ e D / w ( x 1 , t ) ] = 1 h ( x 2 , t ) 1 h ( x 1 , t ) = exp 0 1 d d s log [ 1 h ( ζ ( s ) , t ) ] d s exp [ log α ( R ) ] = α 1 ( R ) ,

and so the claim follows at once by a further exponentiation and rearranging terms.

6 Proof of the local Hamilton estimate in Theorem 3.4

This section is devoted to the proof of Theorem 3.4 and its ingredients. The first subsection develops an evolution inequality on a quantity (below called F β α ) built out of the solution w, where unlike the previous case, there are now two parameters α > 1, β ≥ 0 present, and the last section completes the proof by putting these together, using a localisation argument and finally concluding by an application of the maximum principle.

6.1 Evolution inequality for F β α = w ( β + 2 ) / α 2 | w | 2 / α 2

From the positive solution w we first define a positive function through f = w 1/α (with α > 1 fixed) and then F β α = f β | f | 2 (with β ≥ 0 fixed). The evolution of F = F β α under the equation ( P ) is the subject of the next lemma.

Lemma 6.1.

Let w be a positive solution to ( P ) . For α > 1 and β ≥ 0 fixed constants put f = w 1/α and F β α = f β | f | 2 = w ( β + 2 ) / α 2 | w | 2 / α 2 . Then F = F β α ( x , t ) satisfies the evolution equation

(6.1) [ Δ ϕ t ] F = 2 f β R ic ϕ ( f , f ) + 2 [ 1 α + β ] f β 1 f , | f | 2 + 2 f β | f | 2 [ 2 β 2 α ( 2 β ) ] F 2 / f β + 2 2 f α G w ( t , x , f α ) [ 2 ( 2 + β ) / α ] G ( t , x , f α ) F / f α ( 2 / α ) f β + 1 α f , G x ( t , x , f α ) .

In particular, if R ic ϕ ( g ) ( n 1 ) k g for some k ≥ 0, then

(6.2) [ Δ ϕ t ] F 2 ( n 1 ) k F + 2 ( 1 α ) f , F / f + [ 2 α 2 β 2 β ( 2 α ) ] F 2 / f β + 2 2 f α G w ( t , x , f α ) [ 2 ( 2 + β ) / α ] G ( t , x , f α ) F / f α ( 2 / α ) f β + 1 α f , G x ( t , x , f α ) .

Proof.

It follows from the defining relation f = w 1/α that ∂ t w = αf α−1 t f and Δ ϕ w = αf α−1Δ ϕ f + α(α − 1)f α−2|∇f|2. Thus, from ( P ) it follows that

(6.3) t f = Δ ϕ f + ( α 1 ) | f | 2 / f + ( 1 / α ) f 1 α G ( t , x , f α ) .

Next, an application of the weighted Bochner-Weitzenböck formula and the evolution of f described by (6.3) gives,

(6.4) [ Δ ϕ t ] | f | 2 = 2 | f | 2 + 2 f , Δ ϕ f + 2 R ic ϕ ( f , f ) t | f | 2 = 2 | f | 2 + 2 f , ( Δ ϕ f t f ) + 2 R ic ϕ ( f , f ) = 2 | f | 2 + 2 ( 1 α ) f , [ | f | 2 / f ] ( 2 / α ) f , [ f 1 α G ( t , x , f α ) ] + 2 R ic ϕ ( f , f ) .

In a similar way we have

(6.5) [ Δ ϕ t ] f β = β f β 1 [ Δ ϕ t ] f + β ( β 1 ) f β 2 | f | 2 = β f β 1 [ ( 1 α ) | f | 2 / f ( 1 / α ) f 1 α G ( t , x , f α ) ] + β ( β 1 ) f β 2 | f | 2 = β ( β α ) f β 2 | f | 2 ( β / α ) f β α G ( t , x , f α ) .

As by a straightforward calculation

(6.6) [ Δ ϕ t ] ( f β | f | 2 ) = f β [ Δ ϕ t ] | f | 2 + 2 f β , | f | 2 + | f | 2 [ Δ ϕ t ] f β ,

it then follows by putting (6.4)(6.6) together and rearranging the equation by moving the terms involving the nonlinearity G to the end that

(6.7) [ Δ ϕ t ] F = 2 f β | f | 2 + 2 ( 1 α ) f β 1 [ f , | f | 2 | f | 4 / f ] + 2 f β R ic ϕ ( f , f ) + 2 β f β 1 f , | f | 2 + β ( β α ) f β 2 | f | 4 ( β / α ) f β α | f | 2 G ( t , x , f α ) ( 2 / α ) f β f , [ f 1 α G ( t , x , f α ) ] .

Next, by virtue of G ( t , x , f α ) = G x ( t , x , f α ) + α f α 1 G w ( t , x , f α ) f we can compute

(6.8) f , [ f 1 α G ( t , x , f α ) ] = ( 1 α ) | f | 2 G ( t , x , f α ) / f α + f 1 α f , G ( t , x , f α ) = ( 1 α ) | f | 2 G ( t , x , f α ) / f α + α | f | 2 G w ( t , x , f α ) + f 1 α f , G x ( t , x , f α ) .

Therefore by substituting (6.8) back into the (6.7), rearranging terms and recalling the relation F = f β |∇f|2 we can write

(6.9) [ Δ ϕ t ] F = 2 f β | f | 2 + 2 ( 1 α ) f β 1 f , | f | 2 + 2 f β R ic ϕ ( f , f ) + 2 β f β 1 f , | f | 2 [ 2 β 2 α ( 2 β ) ] f β 2 | f | 4 ( β / α ) f β α | f | 2 G ( t , x , f α ) 2 f β | f | 2 G w ( t , x , f α ) ( 2 / α ) f β + 1 α f , G x ( t , x , f α ) 2 [ ( 1 α ) / α ] f β α | f | 2 × G ( t , x , f α ) = 2 f β | f | 2 + 2 ( 1 α ) f β 1 f , | f | 2 + 2 f β R ic ϕ ( f , f ) + 2 β f β 1 f , | f | 2 [ 2 β 2 α ( 2 β ) ] F 2 / f β + 2 2 f α G w ( t , x , f α ) [ 2 ( 2 + β ) / α ] G ( t , x , f α ) F / f α ( 2 / α ) f β + 1 α f , G x ( t , x , f α ) ,

which is the required conclusion as stated in (6.1).

To justify the inequality in the second part we make use of (6.1) together with the Ricci curvature lower bound R ic ϕ ( g ) ( n 1 ) k g and the basic inequaltiy

f β | f | 2 + β f β 1 f , | f | 2 + β 2 f β 2 | f | 4 0 ,

resulting in turn from “completing the square” ideneity

f | f | 2 + β f , | f | 2 = | f f + β [ f f ] / f | 2 β 2 | f | 4 / f .

Therefore substituting in (6.1) and again recalling F = f β |∇f|2 it follows that

(6.10) [ Δ ϕ t ] F 2 ( n 1 ) k f β | f | 2 + 2 ( 1 α ) f β 1 f , | f | 2 [ 2 β 2 α ( 2 β ) ] F 2 / f β + 2 2 β 2 f β 2 | f | 4 2 f α G w ( t , x , f α ) [ 2 ( 2 + β ) / α ] G ( t , x , f α ) F / f α ( 2 / α ) f β + 1 α f , G x ( t , x , f α ) 2 ( n 1 ) k F + 2 ( 1 α ) f β 1 f , | f | 2 [ 2 + β 2 α ( 2 β ) ] F 2 / f β + 2 ( 2 / α ) f β + 1 α f , G x ( t , x , f α ) 2 f α G w ( t , x , f α ) [ 2 ( 2 + β ) / α ] G ( t , x , f α ) F / f α .

By making note of ⟨∇f, ∇F⟩ = ⟨∇f, ∇(f β |∇f|2)⟩ = βf β−1|∇f|4 + f β ⟨∇f, ∇|∇f|2⟩ and substituting back in (6.10) for F we thus conclude

(6.11) [ Δ ϕ t ] F 2 ( n 1 ) k F + 2 ( 1 α ) f , F / f 2 β ( 1 α ) F 2 / f β + 2 [ 2 + β 2 α ( 2 β ) ] F 2 / f β + 2 ( 2 / α ) f β + 1 α f , G x ( t , x , f α ) 2 f α G w ( t , x , f α ) [ 2 ( 2 + β ) / α ] G ( t , x , f α ) F / f α 2 ( n 1 ) k F + 2 ( 1 α ) f , F / f + [ 2 α 2 β 2 β ( 2 α ) ] × F 2 / f β + 2 ( 2 / α ) f β + 1 α f , G x ( t , x , f α ) 2 f α G w ( t , x , f α ) [ 2 ( 2 + β ) / α ] G ( t , x , f α ) F / f α .

which is the required conclusion as stated in (6.2). This completes the proof. □

6.2 Finalising the proof of Theorem 3.4

Let us now turn on to completing the proof of the local estimate in Theorem 3.4. Here, for reasons that will become clear shortly, the ranges of the parameters α, β will be restricted to β ≥ 0, α > 1 + β (see the last stage of the proof). Towards this end recalling the inequality (6.2) in Lemma 6.1 we have

(6.12) [ Δ ϕ t ] F 2 ( 1 α ) f , F / f + [ 2 α 2 β 2 β ( 2 α ) ] F 2 / f β + 2 2 ( n 1 ) k F ( 2 / α ) f β + 1 α f , G x ( t , x , f α ) 2 f α G w ( t , x , f α ) [ 2 ( 2 + β ) / α ] G ( t , x , f α ) F / f α .

Next, localising by taking a space-time cut-off function η as in (4.11) and following similar principles to those used in the proof of Theorem 3.1, we can write

(6.13) [ Δ ϕ t ] ( η F ) 2 ( 1 α ) f / f + η / η , ( η F ) 2 F ( 1 α ) f / f + η / η , η + [ 2 α 2 β 2 β ( 2 α ) ] η F 2 / f β + 2 + F [ Δ ϕ t 2 ( n 1 ) k ] η ( 2 / α ) f β + 1 α η f , G x ( t , x , f α ) 2 f α G w ( t , x , f α ) [ 2 ( 2 + β ) / α ] G ( t , x , f α ) η F / f α .

For fixed τ ∈ (t 0T, t 0] let (x 1, t 1) be a maximum point for the localised function ηF in the compact set {(x, t): d(x, x 0) ≤ R, t 0Ttτ}. Then without loss of generality we can take t 1 > t 0T and for the sake of establishing the estimate at (x, τ) in Q R/2,T it suffices to confine to the case d(x 1, x 0) ≥ 1. Now at (x 1, t 1) we have the inequalities

(6.14) t ( η F ) 0 , ( η F ) = 0 , Δ ( η F ) 0 .

Therefore applying these to (6.13) and rearranging the inequality result in

(6.15) [ 2 α 2 β 2 β ( 2 α ) ] η F 2 2 f β + 2 F [ ( 1 α ) f / f + η / η ] , η f β + 2 F [ Δ ϕ t 2 ( n 1 ) k ] η + ( 2 / α ) f 2 β + 3 α η f , G x ( t , x , f α ) + 2 f α G w ( t , x , f α ) [ 2 ( 2 + β ) / α ] G ( t , x , f α ) f β + 2 α η F .

We now proceed onto bounding from above each of the terms on the right-hand side of (6.15). Again the argument proceeds by considering the two case d(x 1, x 0) ≤ 1 and d(x 1, x 0) ≥ 1. However in view of certain similarities with the proof of Theorem 3.1, we shall remain brief, focusing on case two only and mainly on the differences.

  1. First the sum on the first line on the right-hand side of (6.15) is bounded directly in modulus by using the Cauchy-Schwarz followed by Young’s inequality:

    (6.16) 2 ( 1 α ) f β + 1 F f , η 2 ( 1 α ) f β + 1 F f , η 2 ( α 1 ) η 3 / 4 F 3 / 2 f 1 + β / 2 | η | η 3 / 4 ( α 1 ) 4 η F 2 + C ( α ) | η | 4 η 3 f 2 β + 4 ( α 1 ) 4 η F 2 + C ( α ) R 4 f 2 β + 4 ,

    where we have used F = f β / 2 | f | and in much the same way

    (6.17) 2 | η | 2 η f β + 2 F = 2 η 1 / 2 F | η | 2 η 3 / 2 f β + 2 α 1 4 η F 2 + C ( α ) | η | 4 η 3 f 2 β + 4 α 1 4 η F 2 + C ( α ) R 4 f 2 β + 4 .

  2. For the terms involving Δ ϕ η and ∂ t η as in the proof of Theorem 3.1 we have the bounds

    Δ ϕ η C 1 R 2 + [ γ Δ ϕ ] + R + ( n 1 ) k η and  t η c η τ t 0 + T .

    Hence, by substituting and making use of Young’s inequality we have,

    (6.18) f β + 2 F Δ ϕ t 2 ( n 1 ) k η C ( n ) F η 1 τ t 0 + T + k + 1 R 2 + [ γ Δ ϕ ] + R f β + 2 α 1 4 η F 2 + C ( α , n ) 1 ( τ t 0 + T ) 2 + k 2 + 1 R 4 + [ γ Δ ϕ ] + 2 R 2 f 2 β + 4 .

  3. Making note of | f , G x ( t , x , f α ) | | f | | G x ( t , x , f α ) | and w = f α we have by abbreviating the arguments (t, x, f α )

    (6.19) 2 w G w [ 2 ( 2 + β ) / α ] G w f β + 2 η F + ( 2 / α ) f 2 β + 3 f , G x η w 2 w G w [ 2 ( 2 + β ) / α ] G w + f β + 2 η F + ( 2 / α ) f 3 β / 2 + 3 F | G x | w η α 1 4 η F 2 + C ( α ) f 2 β + 4 | G x ( t , x , w ) | w 4 / 3 + 2 w G w ( t , x , w ) [ 2 ( 2 + β ) / α ] G ( t , x , w ) w + 2 ,

Now putting together (6.16)(6.19) it follows from (6.15) after some basic calculations and rearrangement of terms that at the space-time point (x 1, t 1) we have

(6.20) [ α 1 β 2 β ( 2 α ) ] η F 2 C ( α , n ) sup Q R , T w ( 2 β + 4 ) / α × k 2 + 1 ( τ t 0 + T ) 2 + 1 R 4 + [ γ Δ ϕ ] + 2 R 2 + sup Q R , T | G x ( t , x , w ) | w 4 / 3 + sup Q R , T 2 w G w ( t , x , w ) [ 2 ( 2 + β ) / α ] G ( t , x , w ) w + 2 .

By the maximality of ηF at (x 1, t 1) we have for any x with d(x, x 0) ≤ R/2 the chain of inequalities F ( x , τ ) [ η F ] ( x , τ ) [ η F ] ( x 1 , t 1 ) [ η F ] ( x 1 , t 1 ) [recall that η(x, τ) = 1 when d(x, x 0) ≤ R/2]. Hence combining the latter with (6.20) and making note of the relations f α = w and F = f β |∇f|2 = w (β+2)/α |∇w|2/(α 2 w 2) this gives

s ( α , β ) | w | 2 w 2 ( β + 2 ) / α = s ( α , β ) α 2 f β | f | 2 C ( α , n ) sup Q R , T w ( β + 2 ) / α

(6.21) × k + 1 τ t 0 + T + 1 R 2 + [ γ Δ ϕ ] + R + sup Q R , T | G x ( t , x , w ) | w 2 / 3 + sup Q R , T 2 w G w ( t , x , w ) [ 2 ( 2 + β ) / α ] G ( t , x , w ) w + ,

where we have set

(6.22) s ( α , β ) = [ α 1 β 2 β ( 2 α ) ] 1 / 2 α 2 = [ 1 ( 1 + β 2 ) / α ( 2 / α 1 ) β ] 1 / 2 α 3 / 2 = [ ( 1 + β ) ( 1 ( 1 + β ) / α ) ] 1 / 2 α 3 / 2 .

Note that the condition 0 < (1 + β) < α guarantees that (1 + β)[1 − (1 + β)/α] > 0 and so in turn s(α, β) > 0. This upon rearranging terms, taking square roots and noting the arbitrariness of τ gives the desired estimate for every ( x , t ) B R / 2 × ( t 0 T , t 0 ] .

7 Proof of the Liouville theorems and implications (I): R i c ϕ ( g ) 0

This section is devoted to the proofs of the Liouville constancy results formulated in Theorem 3.6 through to Theorem 3.11. Note that in Theorems 3.63.10 it is the elliptic counterpart of ( P ) with G = G ( w ) that is being considered and so evidently here the solution w is time independent.

Proof of Theorem 3.6.

Let h = log(w/D) and note that in view of 0 < wD we have h ≤ 0 and so 1/(1 − h) ≤ 1. Referring to the estimate (3.1), making note of G = G ( w ) , the solution w being time independent, k = 0 and then passing to the limit t ↗ ∞ (hence 1 / t t 0 + T 0 ) gives for each fixed x in B R / 2 the bound

(7.1) | w | w ( x ) C 1 log w D ( x ) × k + 1 R + [ γ Δ ϕ ] + R + sup B R | G x ( w ) | w [ 1 log ( w / D ) ] 2 1 3 + sup B R w [ 1 log ( w / D ) ] G w ( w ) + log ( w / D ) G ( w ) w [ 1 log ( w / D ) ] 2 + 1 2

C ( 1 h ( x ) ) 1 / R + [ γ Δ ϕ ] + 1 / 2 / R + sup B R [ ( 1 h ) G ( w ) + h G ( w ) / w ] + 1 / 2 .

From ( 1 h ) w G ( w ) + h G ( w ) 0 it follows that [ ( 1 h ) G ( w ) + h G ( w ) / w ] + 0 and passing to the limit R ↗ ∞ gives |∇w|(x) = 0. The arbitrariness of xM now gives |∇w| ≡ 0 on M and the connectedness of M implies that w is a constant. □

Proof of Theorem 3.7.

As the methodology and main ideas here are to some extenet similar to those used in the proof of Theorem 3.6, we shall remain brief, focusing mainly on the differences. Referring to the estimate (3.7), making note of G = G ( w ) , the solution w being time independent, k = 0 and then passing to the limit t ↗ ∞ (hence 1 / t t 0 + T 0 ) we have for each fixed x in B R / 2 the bound

(7.2) | w | w 1 ( β + 2 ) / ( 2 α ) ( x ) C sup B R w ( β + 2 ) / ( 2 α )

× k + 1 R + [ γ Δ ϕ ] + R + sup B R | G x ( w ) | / w 1 3 + sup B R 2 G w ( w ) [ 2 ( β + 2 ) / α ] G ( w ) / w + 1 2 ,

C sup B R w ( β + 2 ) / ( 2 α ) 1 / R + [ γ Δ ϕ ] + 1 / 2 / R + sup B R 2 G ( w ) [ 2 ( β + 2 ) / α ] G ( w ) / w + 1 2 .

Now [ 1 ( β / 2 + 1 ) / α ] G ( w ) w G ( w ) 0 gives [ 2 G ( w ) [ 2 ( β + 2 ) / α ] G ( w ) / w ] + 0 and making note of supw < ∞ and (β + 2)/(2α) > 0 and passing to the limit R ↗ ∞ gives |∇w|(x) = 0. The arbitrariness of xM now gives |∇w| ≡ 0 on M and again the connectedness of M implies that w is a constant. □

Proof of Theorem 3.8.

Consider the function G ( w ) = X ( w ) + w r Y ( log w ) where w > 0. A basic differentiation then gives G ( w ) = X ( w ) + w r 1 [ r Y ( log w ) + Y ( log w ) ] and by so using the sub-additivity of [⋅]+ and recalling that h = log(w/D) we can write

(7.3) L = w ( 1 h ) G ( w ) + h G ( w ) ( 1 h ) 2 w + { w ( 1 h ) X ( w ) + h X ( w ) } ( 1 h ) 2 w + + w r { [ ( 1 h ) r + h ] Y ( log w ) + ( 1 h ) Y ( log w ) } ( 1 h ) 2 w + = L 1 + L 2 .

By inspection it is seen that (1 − h)wX′ + hX = wX′ − h(wX′ − X) ≤ 0 for w > 0 as a result of the assumptions on X in the theorem. Therefore

(7.4) L 1 = ( 1 h ) w X ( w ) + h X ( w ) ( 1 h ) 2 w + = w X ( w ) + h [ X ( w ) w X ( w ) ] ( 1 h ) 2 w + 0 .

Next we have [r + h/(1 − h)]Y(log w) + Y′(log w) ≤ 0 for w ≥ 1, r ≥ 1 when Y satisfies ( Y 1 ) . Thus here we have

(7.5) L 2 = w r { [ ( 1 h ) r + h ] Y ( log w ) + ( 1 h ) Y ( log w ) } ( 1 h ) 2 w + = w r 1 1 h r + h 1 h Y ( log w ) + Y ( log w ) + 0 .

Similarly from ( Y 2 ) we have sY′(s) ≥ γY(s) for s ≤ 0 and so when r ≤ min(γ, 1) and 0 < w ≤ 1 we have

(7.6) r h ( r 1 ) 1 h Y ( log w ) + Y ( log w ) = [ r h ( r 1 ) ] Y ( log w ) + ( 1 h ) Y ( log w ) 1 h [ r γ h ( r 1 ) ] Y ( log w ) 1 h + ( 1 + log D ) Y ( log w ) 1 h 0 ,

by writing h = log(w/D) and choosing D ≥ 1. Therefore again here we have

(7.7) L 2 = w r { [ ( 1 h ) r + h ] Y ( log w ) + ( 1 h ) Y ( log w ) } ( 1 h ) 2 w + = w r 1 1 h r h ( r 1 ) 1 h Y ( log w ) + Y ( log w ) + 0 .

Thus referring to in (7.3) we have L = 0 . An application of the estimate (3.1) as in the proof of Theorem 3.7 now gives the desired conclusion. □

Proof of Theorem 3.9.

The idea is to invoke Theorem 3.7 with the choices X = X(w) as in (3.14) and Y ≡ 0. Indeed referring to the formulation of Theorem 3.7 it is seen by a direct calculation that

(7.8) X ( w ) = j = 1 N p j A j w p j 1 + j = 1 N q j B j w q j 1 ,

and thus

(7.9) [ 1 ( β / 2 + 1 ) / α ] X ( w ) w X ( w ) = j = 1 N A j [ 1 ( β / 2 + 1 ) / α p j ] w p j + j = 1 N B j [ 1 ( β / 2 + 1 ) / α q j ] w q j .

This is then easily seen to be non-negative, as required by Theorem 3.7, upon suitably restricting the ranges of A j , B j and p j , q j as formulated in the statement of the theorem.

Remark 7.1.

Note that the upper and lower bounds on the exponents p j , q j are given by the same quantity 1 − (β/2 + 1)/α which can be adjusted by optimising the parameters α, β within their respective range. In particular if B j w q j 0 we only require A j 0 and p j < 1 (with α = 0, β ↘ 0) and if A j w p j 0 we only require B j 0 and q j > 0 (with α = 0, β ↗ 1).

Proof of Theorem 3.10.

Writing G ( w ) = A w p + B w q + w r Y ( log w ) a straightforward calculation gives

(7.10) [ 1 ( β / 2 + 1 ) / α ] G ( w ) w G ( w ) = w r ( [ 1 ( β / 2 + 1 ) / α r ] Y Y ) + A [ 1 ( β / 2 + 1 ) / α p ] w p + B [ 1 ( β / 2 + 1 ) / α q ] w q .

The conclusion follows from an application of the first part of Theorem 3.7 upon noting that according to assumptions [ 1 ( β / 2 + 1 ) / α ] G ( w ) w G ( w ) 0 . □

Proof of Theorem 3.11.

Fix a space-time point (x 0, t 0). Then with the quantities k = 0, t = t 0 and T = R in (3.1) and upon writing D R , R = sup Q R , T w and making note of log w ( x , t ) = o [ r ( x ) + | t | ] we have

(7.11) | w | w ( x 0 , t 0 ) C 1 log w ( x 0 , t 0 ) + log D R , R × k + 1 t t 0 + T + 1 R + [ γ Δ ϕ ] + R + sup Q R , T | G x ( w ) | w [ 1 log ( w / D ) ] 2 1 3 + sup Q R , T w [ 1 log ( w / D ) ] G w ( w ) + log ( w / D ) G ( w ) w [ 1 log ( w / D ) ] 2 + 1 2

C 1 log w ( x 0 , t 0 ) + log D R , R 1 / R + 1 / R + [ γ Δ ϕ ] + 1 / 2 / R

where as in the proof of Theorem 3.8 we have deduced from the assumptions G ( w ) 0 and G ( w ) w G ( w ) 0 that [ ( 1 log ( w / D ) ) w G ( w ) + log ( w / D ) G ( w ) ] + = 0 . Passing to the limit R ↗ ∞ and using the assumption on the growth of w = w(x, t) it follows that |∇w(x 0, t 0)| = 0. The arbitrariness of (x 0, t 0) implies |∇w| ≡ 0 and so w = w(t). From the Equation (3.19) it then follows that d w / d t = G ( w ) . Integrating the latter and using the assumption G ( w ) a > 0 for all w > 0 gives w(t) ≤ w(0) + at for all t < 0. This however clashes with w(t) > 0 as t ↘ −∞ and so the conclusion is reached. □

8 Proof of the local Li-Yau estimate in Theorem 3.12

We now attend to the proof of the local Li-Yau estimate (also called a differential Harnack inequality) as formulated in Theorem 3.12. The first two subsections develop the necessary tools including an evolution inequality on a suitably defined Harnack quantity (below called F) built out of the solution w, and involving a parameter α > 1, and the last section completes the proof by putting these together, using a localisation argument and eventually maximum principle.

8.1 Evolution of the Harnack quantity F = t | w | 2 / w 2 α t w / w + α G / w

From the positive solution w we first define the function f = log w and then define the Harnack quantity F = t | f | 2 α t f + α e f G ( t , x , e f ) (for all t ≥ 0 and fixed non-zero parameter α). The evolution of F = F α under the equation ( P ) is the subject of the next lemma.

Lemma 8.1.

Let w be a positive solution to ( P ) . For f = log w and α a fixed non-zero constant let F = F α (x, t) be defined by

(8.1) F = t | w | 2 w 2 α t w w + α G ( t , x , w ) w = t | w | 2 w 2 α Δ ϕ w w = t | f | 2 α t f + α e f G ( t , x , e f ) , t 0 .

Then F satisfies the evolution equation

(8.2) ( Δ ϕ t ) F = 2 t | f | 2 2 f , F + 2 t R ic ϕ m ( f , f ) + [ 2 t / ( m n ) ] ϕ , f 2 F / t + 2 t ( α 1 ) f , [ e f G ( t , x , e f ) ] + α t Δ ϕ [ e f G ( t , x , e f ) ] .

In particular, if R ic ϕ m ( g ) ( m 1 ) k g for some nm < ∞ and k ≥ 0 then

(8.3) ( Δ ϕ t ) F 2 t ( Δ ϕ f ) 2 / m 2 f , F 2 t ( m 1 ) k | f | 2 F / t + 2 t ( α 1 ) f , ( e f G ( t , x , e f ) ) + α t Δ ϕ [ e f G ( t , x , e f ) ] .

Proof.

From ( P ) it follows that f = log w satisfies the equation

(8.4) t f = Δ ϕ f + | f | 2 + e f G ( t , x , e f ) .

Thus using (8.1) and (8.4) it is seen that

(8.5) Δ ϕ f = t f | f | 2 e f G ( t , x , e f ) = ( 1 / α ) | f | 2 t f + e f G ( t , x , e f ) [ ( α 1 ) / α ] | f | 2 = F / ( α t ) [ ( α 1 ) / α ] | f | 2 , t > 0 .

Next, an application of Δ ϕ to F in (8.1) gives

(8.6) Δ ϕ F = t ( Δ ϕ | f | 2 α Δ ϕ ( t f ) + α Δ ϕ [ e f G ( t , x , e f ) ] ) = 2 t | f | 2 + 2 t f , Δ ϕ f + 2 t R ic ϕ m ( f , f ) + [ 2 t / ( m n ) ] ϕ , f 2 α t t ( Δ ϕ f ) + α t Δ ϕ [ e f G ( t , x , e f ) ]

where in concluding the second line we have used the weighted Bocnher-Weitzenböck formula as applied to f. Hence by virtue of the (8.4) and (8.5) this can be simplified and re-written as

(8.7) Δ ϕ F = 2 t | f | 2 + 2 t R ic ϕ m ( f , f ) + [ 2 t / ( m n ) ] ϕ , f 2 + 2 t f , Δ ϕ f + ( t t F F ) / t + 2 t ( α 1 ) f , ( t f ) + α t Δ ϕ [ e f G ( t , x , e f ) ] = 2 t | f | 2 + 2 t R ic ϕ m ( f , f ) + [ 2 t / ( m n ) ] ϕ , f 2 + 2 t ( α 1 ) f , [ e f G ( t , x , e f ) ] + t F F / t 2 f , F + α t Δ ϕ [ e f G ( t , x , e f ) ] .

Rearranging this now lead to

(8.8) ( Δ ϕ t ) F = 2 t | f | 2 + 2 t R ic ϕ m ( f , f ) + [ 2 t / ( m n ) ] ϕ , f 2 F / t 2 f , F + 2 t ( α 1 ) f , [ e f G ( t , x , e f ) ] + α t Δ ϕ [ e f G ( t , x , e f ) ] ,

which is the desired conclusion.

For the inequality in (8.3) we first note that by an application of the Cauchy-Schwarz and Young inequalities we have

(8.9) | f | 2 + ϕ , f 2 / ( m n ) ( Δ f ) 2 / n + ϕ , f 2 / ( m n ) ( Δ ϕ f ) 2 / m .

Hence the desired conclusion follows by substituting the above in (8.8) and applying the Ricci curvature lower bound R ic ϕ m ( g ) ( m 1 ) k g as given in the lemma. □

In the course of the proof of Theorem 3.12 we make use of the following lemma. As before we use the notation G x : x G ( t , x , w ) for the function obtained by freezing the variables t, w and viewing G solely as a function of x. We also denote partial derivatives of G with subscripts.

Lemma 8.2.

Suppose G = G ( t , x , w ) is a twice continuously differentiable function on [0, T] × M × (0, ∞) and let w = e f with f = f(x, t) be a twice continuously differentiable on M × [0, T]. Then

(8.10) Δ ϕ G = Δ ϕ G x + 2 e f G x w , f + e f | f | 2 G w + e f G w w + e f G w Δ ϕ f .

Proof.

Firstly, it is seen, by calculating in local coordinates or directly otherwise that with G = G ( t , x , w ) and w = e f we have

(8.11) G = G x + e f G w f , G x = ( G x 1 , , G x n ) ,

where we have abbreviated the arguments (t, x, w) for convenience. Now differentiating further with the aim of calculating the Laplacian (and noting G x = G x ) gives

(8.12) Δ G = Δ G x + e f G x w , f + e f | f | 2 G w + e f G x w , f + e 2 f | f | 2 G w w + e f G w Δ f = Δ G x + 2 e f G x w , f + e f | f | 2 G w + e f G w w + e f G w Δ f .

Subsequently calculating the Witten Laplacian, by utilising the above fragments we have,

(8.13) Δ ϕ G = Δ G ϕ , G = Δ G ϕ , G x + e f G w f = Δ G ϕ , G x e f G w ϕ , f = Δ G x ϕ , G x + 2 e f G x w , f + e f | f | 2 G w + e f G w w + e f G w Δ ϕ f ,

which is the required result. □

8.2 Construction of spatial cut-offs and localisation in space

The next and final ingredient needed in the proof of Theorem 3.12 is the construction of a suitable cut-off function in space. To this end we pick a reference point x 0M and fix R > 0 and with r(x) = d(x, x 0) being the geodesic radial variable with respect to x 0 we set

(8.14) ζ ( x ) = ζ ̄ r ( x ) R .

The function ζ ̄ = ζ ̄ ( s ) appearing in (8.14) is defined on the half-line s ≥ 0 and can be constructed easily with properties described below (see [14], [15], [35]).

Lemma 8.3.

There exists a function ζ ̄ : [ 0 , ) R such that:

  1. ζ ̄ is of class C 2 [ 0 , ) ,

  2. 0 ζ ̄ ( s ) 1 for 0 ≤ s < ∞ with ζ ̄ ( s ) 1 for s ≤ 1 and ζ ̄ ( s ) 0 for s ≥ 2.

  3. ζ ̄ is non-increasing, ( ζ ̄ 0 ) and additionally for suitable constants c 1, c 2 > 0,

    (8.15) c 1 ζ ̄ ζ ̄ 0 and ζ ̄ c 2 .

It is evident that here ζ ≡ 1 for when 0 ≤ r(x) ≤ R and ζ ≡ 0 for when r(x) ≥ 2R. Furthermore a straightforward calculation starting from (8.14) gives

(8.16) ζ = ( ζ ̄ / R ) r , Δ ζ = ζ ̄ | r | 2 / R 2 + ζ ̄ Δ r / R

and so

(8.17) [ Δ ϕ t ] ζ = Δ ϕ ζ = ζ ̄ | r | 2 / R 2 + ζ ̄ Δ ϕ r / R .

8.3 Finalising the proof of Theorem 3.12

We now come to the proof of the main result in the section which is the local Li-Yau estimate in Theorem 3.12. To this end we consider the spatially localised function ζF where F is as in (8.1). We denote by (x 1, t 1) the point where this function attains its maximum over the compact cylinder {d(x, x 0) ≤ 2R, 0 ≤ tT}. We also assume that [ζF](x 1, t 1) > 0 as otherwise the desired estimate is trivially true as a result of F ≤ 0. It thus follows that t 1 > 0 and d(x 1, x 0) < 2R and so at the maximum point (x 1, t 1) we have the relations

(8.18) t ( ζ F ) 0 , ( ζ F ) = 0 , Δ ( ζ F ) 0 .

From the inequalities above we deduce that

(8.19) 0 [ Δ ϕ t ] ( ζ F ) = F [ Δ ϕ t ] ζ + 2 ζ , F + ζ [ Δ ϕ t ] F .

Now making note of [Δ ϕ − ∂ t ]ζ = Δ ϕ ζ (as ∂ t ζ ≡ 0) it follows from the above that at the maximum point (x 1, t 1) we have

(8.20) 0 F Δ ϕ ζ + 2 ζ , F + ζ [ Δ ϕ t ] H F Δ ϕ ζ + ( 2 / ζ ) ζ , ( ζ F ) 2 ( | ζ | 2 / ζ ) F + ζ [ Δ ϕ t ] F F Δ ϕ ζ 2 ( | ζ | 2 / ζ ) F + ζ [ Δ ϕ t ] F .

Next by referring to (8.16), (8.17) and Δ ϕ r ( m 1 ) k coth ( k r ) [the latter being a consequence of the generalised Laplacian comparison theorem and the lower curvature bound R ic ϕ m ( m 1 ) k g ] it follows upon recalling ζ ̄ 0 in (iii) in Lemma 8.3 that

(8.21) Δ ϕ ζ = ζ ̄ | r | 2 R 2 + ζ ̄ Δ ϕ r R 1 R 2 ζ ̄ + ( m 1 ) R ζ ̄ k coth ( k r ) .

Moreover upon noting coth ( k r ) coth ( k R ) and k coth ( k R ) ( 1 + k R ) / R , for Rr ≤ 2R, we deduce that ( m 1 ) ζ ̄ k coth ( k r ) ( m 1 ) ( 1 + k R ) / R ζ ̄ . Hence by putting the above fragments together we have the lower bound on Δ ϕ ζ in terms of k, R and the constants c 1, c 2 in (8.15):

(8.22) Δ ϕ ζ 1 R 2 ζ ̄ + ( m 1 ) R ζ ̄ k coth ( k r ) 1 R 2 ζ ̄ + ( m 1 ) R 1 R + k ζ ̄ c 2 R 2 ( m 1 ) R c 1 1 R + k = 1 R 2 [ c 2 + ( m 1 ) c 1 ( 1 + R k ) ] .

Likewise, referring to (8.15) and (8.16), a straightforward calculation gives,

(8.23) | ζ | 2 ζ = ζ ̄ 2 ζ ̄ | r | 2 R 2 = ζ ̄ ζ ̄ 2 | r | 2 R 2 c 1 2 R 2 .

Now returning to (8.20), invoking (8.3) and making note of (8.22) and (8.23), we obtain, at the maximum point (x 1, t 1), the inequality

(8.24) 0 F Δ ϕ ζ 2 ( | ζ | 2 / ζ ) F + ζ [ Δ ϕ t ] F F ( [ c 2 + ( m 1 ) c 1 ( 1 + R k ) ] / R 2 ) 2 c 1 2 F / R 2 + ζ 2 t 1 ( Δ ϕ f ) 2 / m F / t 1 2 f , F 2 t 1 ( m 1 ) k | f | 2 + 2 t 1 ( α 1 ) f , ( e f G ) + α t 1 Δ ϕ ( e f G ) .

We point out that here and below we abbreviate the arguments of G = G ( t , x , e f ) for convenience. Next, in view of ∇(ζF) = 0 at (x 1, t 1), that ζ⟨∇f, ∇F⟩ = −F⟨∇f, ∇ζ⟩. Substituting these in (8.24) we can write after rearranging terms

(8.25) 0 F c 2 + ( m 1 ) c 1 ( 1 + R k ) + 2 c 1 2 / R 2 ζ F / t 1 + 2 t 1 ζ / m | f | 2 + e f G t f 2 + 2 F f , ζ 2 t 1 ζ ( m 1 ) k | f | 2 + 2 t 1 ζ ( α 1 ) f , ( e f G ) + α t 1 ζ Δ ϕ ( e f G ) .

Now upon utilising the bound 2 F f , ζ 2 F | f | | ζ | 2 F | f | ( c 1 / R ) ζ it follows after multiplying (8.25) through by t 1 ζ(x 1) = t 1 ζ and rearranging terms, that

(8.26) 0 t 1 ζ F ( c 2 + ( m 1 ) c 1 ( 1 + R k ) + 2 c 1 2 / R 2 + ζ / t 1 ) + 2 t 1 2 ζ 2 / m | f | 2 + e f G t f 2 2 t 1 ( c 1 / R ) ζ 3 / 2 | f | F 2 t 1 2 ζ 2 ( m 1 ) k | f | 2 + t 1 2 ζ 2 2 ( α 1 ) f , ( e f G ) + α Δ ϕ ( e f G ) .

Referring next to the expression on the last line in (8.26) by working on the sum inside the brackets we can write

(8.27) 2 ( α 1 ) f , ( e f G ) + α Δ ϕ ( e f G ) = 2 ( α 1 ) [ e f G | f | 2 + e f f , G ] + α e f Δ ϕ G 2 e f f , G + G Δ ϕ e f = 2 ( α 1 ) e f G | f | 2 + 2 α e f f , G 2 e f f , G x + e f G w f + α e f G x x ϕ , G x + 2 e f G x w , f + e f | f | 2 G w + e f G w w + e f G w Δ ϕ f + α G e f Δ ϕ f + | f | 2 2 α e f f , G .

Here we have made note of the relation

(8.28) Δ ϕ e f = Δ e f ϕ , e f = div ( e f f ) + e f ϕ , f = e f Δ f + e f | f | 2 + e f ϕ , f = e f Δ ϕ f | f | 2 .

Next, since according to (8.5) we have,

(8.29) Δ ϕ f α G w α G e f = F α t 1 α 1 α | f | 2 α G w G e f = F G w G e f / t 1 ( α 1 ) | f | 2 G w G e f ,

upon substituting this back in (8.27) it follows that

2 ( α 1 ) f , ( e f G ) + α Δ ϕ ( e f G ) = 2 ( α 1 ) e f G | f | 2 2 e f f , G x 2 G w | f | 2 + α e f ( G x x ϕ , G x ) + 2 α G x w , f + α | f | 2 G w + α | f | 2 e f G w w + α G e f | f | 2 F G w G e f / t 1 ( α 1 ) | f | 2 G w + G e f = | f | 2 2 ( α 1 ) e f G 2 G w + α G w + α e f G w w + α e f G ( α 1 ) G w + ( α 1 ) G e f 2 f , e f G x α G x w + F G w G e f / t 1 α e f Δ ϕ G x .

Therefore, by taking into account the relevant cancellations, after simplifying terms and using basic inequalities, we can write

(8.30) 2 ( α 1 ) f , ( e f G ) + α Δ ϕ ( e f G ) | f | 2 e f G G w + α e f G w w F G w G e f / t 1 2 | f | | e f G x α G x w | + α e f Δ ϕ G x .

As a result making use of the relations (8.11)(8.13) and the inequality (8.30) above and substituting all back into (8.26) and recalling 0 ≤ ζ ≤ 1 we obtain:

(8.31) 0 ζ F ( c 2 + ( m 1 ) c 1 ( 1 + R k ) + 2 c 1 2 t 1 / R 2 + 1 ) t 1 ζ 2 F G w G e f + 2 t 1 2 ζ 2 / m | f | 2 + e f G t f 2 2 c 1 t 1 ζ 3 / 2 | f | F / R + t 1 2 ζ 2 | f | 2 e f G G w + α e f G w w 2 ( m 1 ) k 2 t 1 2 ζ 2 | f | | e f G x α G x w | + α t 1 2 ζ 2 e f Δ ϕ G x .

In order to obtain the desired bounds out of this it is more efficient to introduce the quantities y, z and a , b by setting

(8.32) y = ζ | f | 2 , z = ζ t f e f G , y α z = ζ F / t 1 > 0 , a = ( m 1 ) k + ( 1 / 2 ) sup e f G + G w α e f G w w + : ( x , t ) H 2 R , T , b = sup | e f G x α G x w | : ( x , t ) H 2 R , T .

Now substituting the quantities (8.32) back in (8.31) and recalling 0 ≤ ζ ≤ 1, it follows that

(8.33) 0 ζ F ( c 2 + ( m 1 ) c 1 ( 1 + R k ) + 2 c 1 2 t 1 / R 2 + 1 ) + 2 t 1 2 / m ( y z ) 2 m c 1 / R y ( y α z ) m a y m b y t 1 ζ F G w e f G + + α t 1 2 ζ 2 e f Δ ϕ G x .

In order to proceed further we now state and prove the following lemma.

Lemma 8.4.

Suppose z R , c, y > 0 and α > 1 are arbitrary constants such that yαz > 0. Then for any ɛ ∈ (0, 1) we have

(8.34) ( y z ) 2 ( m c 1 / R ) y ( y α z ) m a y m b y ( y α z ) 2 / α 2 ( m c 1 / R ) 2 α 2 ( y α z ) / [ 8 ( α 1 ) ] ( α 2 m 2 a 2 ) / [ 4 ( 1 ε ) ( α 1 ) 2 ] ( 3 / 4 ) [ m 4 b 4 α 2 / ( 4 ε ( α 1 ) 2 ) ] 1 / 3 .

Proof.

Starting from the expression on the left-hand side in (8.34) we can write for any δ, ɛ by basic considerations

(8.35) ( y z ) 2 ( m c 1 / R ) y ( y α z ) m a y m b y = ( 1 ε δ ) y 2 ( 2 ε α ) y z + z 2 + ( ε y ( m c 1 / R ) y ) ( y α z ) + δ y 2 m a y m b y = ( 1 / α ε / 2 ) ( y α z ) 2 + ( 1 ε δ 1 / α + ε / 2 ) y 2 + ( 1 α + ε α 2 / 2 ) z 2 + ( ε y ( m c 1 / R ) y ) ( y α z ) + δ y 2 m a y m b y .

In particular setting δ = (1/α − 1)2 and ɛ = 2 − 2/α − 2(1/α − 1)2 = 2(α − 1)/α 2 gives 1 − ɛδ − 1/α + ɛ/2 = 0 and 1 − α + ɛα 2/2 = 0 and so by making note of the inequality ε y ( m c 1 / R ) y ( m c 1 / R ) 2 / ( 4 ε ) with ɛ = 2(α − 1)/α 2 > 0 we can deduce from (8.35) that

(8.36) ( y z ) 2 ( m c 1 / R ) y ( y α z ) m a y m b y ( y α z ) 2 / α 2 m 2 c 1 2 α 2 ( y α z ) / [ 8 R 2 ( α 1 ) ] + ( α 1 ) 2 y 2 / α 2 m a y m b y .

Next, considering the last three terms only we can write, for any ɛ ∈ (0, 1),

(8.37) ( α 1 ) 2 y 2 / α 2 m a y m b y ( α 1 ) 2 y 2 / α 2 ( 1 ε ) ( α 1 ) 2 y 2 / α 2 ( α 2 m 2 a 2 ) / [ 4 ( 1 ε ) ( α 1 ) 2 ] m b y ε ( α 1 ) 2 y 2 / α 2 ( α 2 m 2 a 2 ) / [ 4 ( 1 ε ) ( α 1 ) 2 ] m b y ( α 2 m 2 a 2 ) / [ 4 ( 1 ε ) ( α 1 ) 2 ] ( 3 / 4 ) [ m 4 b 4 α 2 / ( 4 ε ( α 1 ) 2 ) ] 1 / 3

where above we have made use of (1 − ɛ)(α − 1)2 y 2/α 2by ≥−(α 2 b 2)/[4(1 − ɛ)(α − 1)2] and ε ( α 1 ) 2 y 2 / α 2 c y ( 3 / 4 ) c 4 / 3 [ α 2 / ( 4 ε ( α 1 ) 2 ) ] 1 / 3 to deduce the first and last inequalities respectively. Substituting back in (8.36) gives the desired inequality. □

Now returning to the inequality (8.33) and making use of Lemma 8.4 it follows that

(8.38) 0 ζ F ( c 2 + ( m 1 ) c 1 ( 1 + R k ) + 2 c 1 2 t 1 / R 2 + 1 ) + 2 t 1 2 / m ( ζ F ) 2 / t 1 2 α 2 m 2 c 1 2 α 2 ( ζ F ) / ( 8 ( α 1 ) R 2 t 1 ) m t 1 2 α 2 a 2 / [ 2 ( 1 ϵ ) ( α 1 ) 2 ] 3 t 1 2 / 2 ( m α 2 b 4 / [ 4 ϵ ( α 1 ) 2 ] ) 1 / 3 t 1 ζ F G w e f G + + α t 1 2 ζ 2 e f Δ ϕ G x .

We now aim to rewrite (8.38) as an inequality involving a quadratic expression in powers of the localised function ζF. To this end upon setting

(8.39) d = c 2 + ( m 1 ) c 1 ( 1 + R k ) + 2 c 1 2 t 1 / R 2 + 1 + m t 1 c 1 2 α 2 / [ 4 ( α 1 ) R 2 ] + t 1 γ C G ( 2 R ) ,

and

(8.40) e = m α 2 a 2 / [ 2 ( 1 ε ) ( α 1 ) 2 ] + ( 3 / 2 ) [ m α 2 b 4 / ( 4 ε ( α 1 ) 2 ) ] 1 / 3 + α γ D G ( 2 R ) ,

where

(8.41) γ C G ( 2 R ) = sup H 2 R , T G w e f G + , γ D G ( 2 R ) = sup H 2 R , T e f Δ ϕ G x + ,

it is seen that (8.38) can be expressed as

(8.42) 0 2 ( ζ F ) 2 / ( m α 2 ) ( ζ F ) d t 1 2 e .

Now basic considerations using (8.42) on quadratics result in

(8.43) ζ F ( m α 2 / 4 ) d + d 2 + 8 t 1 2 e / ( m α 2 ) ( m α 2 / 4 ) 2 d + 8 t 1 2 e / ( m α 2 ) = m α 2 d / 2 + t 1 α m e / 2 .

Since ζ ≡ 1 for d(x, x 0) ≤ R and (x 1, t 1) is the point where ζF attains its maximum on d(x, x 0) ≤ 2R we have

(8.44) F ( x , τ ) = [ ζ F ] ( x , τ ) [ ζ F ] ( x 1 , t 1 ) m α 2 d / 2 + t 1 α m e / 2 .

Therefore recalling (8.1), substituting for d and e from (8.39) and (8.40) above and making noting t 1τ, we can write after dividing both sides ατ,

(8.45) α 1 | f | 2 t f + e f G ( m α / 2 τ ) d + m e / 2 ( m α ) c 2 + ( m 1 ) c 1 ( 1 + R k ) + 2 c 1 2 / 2 R 2 + ( m α / 2 τ ) + ( m α / 2 ) ( γ C G ( 2 R ) + m c 1 2 α 2 / [ 4 ( α 1 ) R 2 ] ) + m / 2 m α 2 a 2 / [ 2 ( 1 ε ) ( α 1 ) 2 ] + ( 3 / 2 ) [ m α 2 b 4 / ( 4 ε ( α 1 ) 2 ) ] 1 / 3 + α γ D G ( 2 R ) 1 / 2 .

Finally using the arbitrariness of 0 < τT it follows after reverting back to w upon noting the relation f = log w and rearranging terms that

(8.46) | w | 2 α w 2 t w w + G w ( m α / 2 ) 1 / t + γ C G ( 2 R ) + ( m α / 2 ) m c 1 2 α 2 / [ 4 ( α 1 ) ] + c 2 + ( m 1 ) c 1 ( 1 + R k ) + 2 c 1 2 / R 2 + m / 2 m α 2 a 2 / [ 2 ( 1 ϵ ) ( α 1 ) 2 ] + ( 3 / 2 ) [ m α 2 b 4 / ( 4 ε ( α 1 ) 2 ) ] 1 / 3 + α γ D G ( 2 R ) 1 / 2 .

Using (3.20) and (3.21) in conjunction with the bounds (3.24), (3.25) and making note of (8.32) it is seen that

(8.47) a = ( m 1 ) k + γ A G , α ( 2 R ) / 2 , b = γ B G , α ( 2 R )

which upon substitution in (8.46) gives the desired estimate as formulated in (3.28).

9 Proof of the parabolic Harnack inequality in Theorem 3.14

This will be shown as a consequence of the local estimate in Theorem 3.12. Here the main idea is to integrate (3.28) along suitable space-times curves described below. Towards this end and upon referring to the expression on the right-hand side of (3.28) let us introduce the constant

(9.1) H = γ E G ( R ) m α 2 R 2 m c 1 2 α 2 4 ( α 1 ) + c 2 + ( m 1 ) c 1 ( 1 + R k ) + 2 c 1 2 ( m α / 2 ) γ C G ( 2 R ) m α 2 m α ( m 1 ) k + γ A G , α ( 2 R ) / 2 2 2 ( 1 ε ) ( α 1 ) 2 + 3 3 m γ B G , α ( 2 R ) 4 2 5 ε α ( α 1 ) 2 1 / 3 + γ D G ( 2 R ) 1 / 2 ,

where in analogy with the other γ-quantities occurring earlier, the first term on the right stands for

(9.2) γ E G ( R ) = inf H R , T w 1 G ( t , x , w ) .

It then follows from Theorem 3.12 that in the space-time cylinder H R,T we have the inequality

(9.3) t w w | w | 2 α w 2 m α 2 t + H .

Note that in (9.1) we can replace γ E G ( R ) with the smaller quantity γ E G ( 2 R ) in order to maintain uniformity in notation (specifically, have all the γ-quantities in the formula defining H with argument 2R). However, as far as the bound (9.3) is concerned, the larger quantity γ E G ( R ) is sufficient and clearly more accurate. Next suppose ζ C 1 ( [ t 1 , t 2 ] ; M ) is an arbitrary curve in M lying entirely in B R and satisfying ζ(t 1) = x 1 and ζ(t 2) = x 2. In particular (ζ(t), t) ∈ H R,T for all t 1tt 2. Differentiating the function log w(ζ(t), t) and using (9.3) it is then seen that

(9.4) d / d t [ log w ( ζ ( t ) , t ) ] = w / w , ζ ̇ ( t ) + t w / w w / w , ζ ̇ ( t ) + | w | 2 / ( α w 2 ) ( m α ) / ( 2 t ) + H = α 1 | w / w + α ζ ̇ ( t ) / 2 | 2 α | ζ ̇ ( t ) | 2 / 4 ( m α ) / ( 2 t ) + H α | ζ ̇ ( t ) | 2 / 4 ( m α ) / ( 2 t ) + H .

Therefore integrating the above inequality gives

(9.5) log w ( x 2 , t 2 ) w ( x 1 , t 1 ) = log w ( ζ ( t ) , t ) t 1 t 2 = t 1 t 2 d d t log w ( ζ ( t ) , t ) d t t 1 t 2 α 4 | ζ ̇ ( t ) | 2 d t t 1 t 2 m α 2 t d t + t 1 t 2 H d t = ( α / 4 ) t 1 t 2 | ζ ̇ ( t ) | 2 d t m α / 2 log ( t 2 / t 1 ) + ( t 2 t 1 ) H .

Hence upon exponentiating we have

(9.6) w ( x 2 , t 2 ) w ( x 1 , t 1 ) exp t 1 t 2 α 4 | ζ ̇ ( t ) | 2 d t exp [ ( t 2 t 1 ) H ] t 2 t 1 m α / 2 ,

or after rearranging terms and rescaling the integral:

(9.7) w ( x 2 , t 2 ) w ( x 1 , t 1 ) exp [ ( t 2 t 1 ) H α L ( x 1 , x 2 , t 2 t 1 ) ] t 2 t 1 m α / 2

where

(9.8) L ( x 1 , x 2 , t 2 t 1 ) = inf ζ Γ 1 4 ( t 2 t 1 ) 0 1 | ζ ̇ ( t ) | 2 d t .

This gives the parabolic Harnack inequality in its local form. Now if the bounds are global as in Theorem 3.13 then by passing to the limit R → ∞ we obtain the global counterpart of the inequality with the constant H adjusted.

10 Proof of the Liouville theorems and implications (II): R i c ϕ m ( g ) 0

In this section we give the proofs of the Liouville-type results formulated in Theorem 3.16 and its implications as given in Theorems 3.17 and 3.18. The basis for all these proofs is the elliptic estimate in Theorem 3.15 below that in term is a consequence of the parabolic Li-Yau estimates in Theorems 3.12 and 3.13. Towards this end we begin by presenting the proof of the estimate (3.33). We emphasise that in all these results we are assuming the curvature condition R ic ϕ m ( g ) 0 with nm < ∞.

Proof of Theorem 3.15.

The proof of (3.33) for the elliptic equation Δ ϕ w + G ( w ) = 0 follows directly from the global estimate (3.30) in Theorem 3.13 by passing to the limit t ↗ ∞, upon noting that w is time independent. Moreover, here, G is independent of x and so by referring to (3.21) and (3.23) it follows that

(10.1) B G α ( w ) 0 , D G ( w ) 0 ,

thus giving γ B G , α = 0 and γ D G = 0 . As a result the middle expression “ M ” on the right in (3.30) reduces to

(10.2) M = m α 2 m α ( m 1 ) k + γ A G , α / 2 2 2 ( 1 ε ) ( α 1 ) 2 + 3 3 m γ B G , α 4 2 5 ε α ( α 1 ) 2 1 / 3 + γ D G 1 / 2 = m α 2 m α ( m 1 ) k + γ A G , α / 2 2 2 ( 1 ε ) ( α 1 ) 2 1 / 2 = m α 4 2 ( m 1 ) k + γ A G , α ( α 1 ) 1 ε .

Putting the above fragments together gives the desired estimate. □

Proof of Theorem 3.16.

Since here R ic ϕ m ( g ) 0 , by taking k = 0 in (3.33) it follows that under the assumptions of the theorem w satisfies the global estimate on M:

(10.3) | w | 2 α w 2 + G ( w ) w m α 4 ( α 1 ) 1 ε γ A G , α + m α 2 γ C G .

Next, since by assumption G ( w ) w G w ( w ) 0 and G ( w ) w G w ( w ) + α w 2 G w w ( w ) 0 , for some α > 1, it follows upon recalling (3.20) and (3.22) that

(10.4) A G α ( w ) = α w G w w ( w ) + G w ( w ) w 1 G ( w ) + 0 , C G α ( w ) = G w ( w ) w 1 G ( w ) + 0 ,

thus giving γ A G , α = 0 and γ C G = 0 . Substituting these back in (10.3) then leads to

(10.5) | w | 2 α w 2 + G ( w ) w 0 .

Since G ( w ) 0 we thus infer that | w | 2 / ( α w 2 ) + G ( w ) / w 0 and therefore |∇w| ≡ 0 on M. The conclusion on w being a constant is now immediate. □

Proof of Theorem 3.17.

A direct calculation gives

(10.6) G ( w ) w G w ( w ) = j = 1 N A j ( 1 p j ) w p j ,

(10.7) G ( w ) w G w ( w ) + α w 2 G w w ( w ) = j = 1 N [ α A j p j ( p j 1 ) A j p j + A j ] w p j = j = 1 N [ A j ( p j 1 ) ( α p j 1 ) ] w p j .

Now if A j 0 we have G ( w ) 0 and the assumption p j ≤ 1 gives G ( w ) w G w ( w ) 0 and G ( w ) w G w ( w ) + α w 2 G w w ( w ) 0 (by choosing α > 1 suitably). □

Proof of Theorem 3.18.

Here G ( w ) w G w ( w ) = B ( 1 s ) w s A w Y ( log w ) and G ( w ) w G w ( w ) + α w 2 G w w ( w ) = A [ ( α 1 ) w Y + α w Y ] + B ( s 1 ) ( α s 1 ) w s . Therefore the conclusion follows from an application of Theorem 3.16. □


Corresponding author: Ali Taheri, School of Mathematical and Physical Sciences, University of Sussex, Falmer, Brighton, UK, E-mail:

Acknowledgments

The authors wish to thank the anonymous reviewers for a careful reading of the paper and useful comments.

  1. Research ethics: Not applicable.

  2. Author contributions: Both authors have equal contribution in this research, have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Competing interests: Authors state no conflict of interest.

  4. Research funding: The authors gratefully acknowledge financial support from the Engineering and Physical Sciences Research Council (EPSRC) through the research grant EP/V027115/1.

  5. Data availability: All data in this research is provided in full in the results section. Additional data is in the public domain at locations cited in the reference section.

References

[1] E. Acerbi and R. Mingione, “Gradient estimates for a class of parabolic systems,” Duke Math. J., vol. 136, pp. 285–320, 2007, https://doi.org/10.1215/s0012-7094-07-13623-8.Search in Google Scholar

[2] T. Aubin, Nonlinear Analysis on Manifolds, New York, Springer, 1982.Search in Google Scholar

[3] M. Bǎileşteanu, X. Cao, and A. Pulemotov, “Gradient estimates for the heat equation under the Ricci flow,” J. Funct. Anal., vol. 258, pp. 3517–3542, 2010, https://doi.org/10.1016/j.jfa.2009.12.003.Search in Google Scholar

[4] M. F. Biduat-Vèron and L. Veròn, “Nonlinear elliptic equations on compact Riemannian manifolds and asymptotics of the Emden equations,” Invent. Math., vol. 106, pp. 489–539, 1991, https://doi.org/10.1007/bf01243922.Search in Google Scholar

[5] V. Bögelein, F. Duzaar, and G. Mingione, “The regularity of general parabolic systems with degenerate diffusion,” Mem. Am. Math. Soc., vol. 221, no. 1041, p. vi+143, 2013. https://doi.org/10.1090/s0065-9266-2012-00664-2.Search in Google Scholar

[6] E. Calabi, “An extension of Hopf maximum principle with application to Riemannian geometry,” Duke Math. J., vol. 25, pp. 45–56, 1958, https://doi.org/10.1215/s0012-7094-58-02505-5.Search in Google Scholar

[7] H. D. Cao, Recent Progress on Ricci solitons, Recent Advances in Geometric Analysis, Advanced Lectures in Mathematics (ALM), vol. 11, International Press, 2010, pp. 1–38.Search in Google Scholar

[8] S. Y. Cheng and S. T. Yau, “Differential equations on Riemannian manifolds and their geometric applications,” Commun. Pure Appl. Math., vol. 28, pp. 333–354, 1975, https://doi.org/10.1002/cpa.3160280303.Search in Google Scholar

[9] B. Chow, P. Lu, and L. Nei, Hamilton’s Ricci Flow, Graduate Studies in Mathematics, vol. 77, AMS, 2006.10.1090/gsm/077Search in Google Scholar

[10] M. Giaquinta, Multiple Integrals in the Calculus of Variations and Nonlinear Elliptic Systems, Annals of Mathematics Studies, vol. 105, Princeton University Press, 1983.10.1515/9781400881628Search in Google Scholar

[11] B. Gidas and J. Spruck, “Global and local behaviour of positive solutions of nonlinear elliptic equations,” Commun. Pure Appl. Math., vol. 34, pp. 525–598, 1981, https://doi.org/10.1002/cpa.3160340406.Search in Google Scholar

[12] A. Grigor’yan, Heat Kernel Analysis on Manifolds, Studies in Advanced Mathematics, AMS, 2013.10.1090/amsip/047Search in Google Scholar

[13] R. Hamilton, “A matrix Harnack estimate for heat equation,” Commun. Anal. Geom, vol. 1, pp. 113–126, 1993, https://doi.org/10.4310/cag.1993.v1.n1.a6.Search in Google Scholar

[14] P. Li, Geometric Analysis, Cambridge Studies in Advanced Mathematics, vol. 134, CUP, 2012.Search in Google Scholar

[15] P. Li and S. T. Yau, “On the parabolic kernel of Schrödinger operator,” Acta Math., vol. 156, pp. 153–201, 1986, https://doi.org/10.1007/bf02399203.Search in Google Scholar

[16] S. Li and X. D. Li, “W-Entropy, super Perelman Ricci flows and (K, m)-Ricci solitons,” J. Geom. Anal., vol. 30, pp. 3149–3180, 2020, https://doi.org/10.1007/s12220-019-00193-4.Search in Google Scholar

[17] S. Li and X. D. Li, “Hamilton’s differential Harnack inequality and the W$\mathcal{W}$-entropy formula on complete Riemannian manifolds,” J. Funct. Anal., vol. 274, pp. 3263–3290, 2018, https://doi.org/10.1016/j.jfa.2017.09.017.Search in Google Scholar

[18] S. Li and X. D. Li, “Harnack inequalities for witten laplacian on riemannian manifolds with super Ricci flows,” Asian J. Math., vol. 22, pp. 577–598, 2018, https://doi.org/10.4310/ajm.2018.v22.n3.a10.Search in Google Scholar

[19] X. D. Li, “Liouville theorems for symmetric diffusion operators on complete Riemannian manifolds,” J. Math. Pure Appl., vol. 84, pp. 1295–1361, 2005, https://doi.org/10.1016/j.matpur.2005.04.002.Search in Google Scholar

[20] G. Morrison and A. Taheri, “An infinite scale of incompressible twisting solutions to the nonlinear elliptic system L[u;A,B]=∇P$\mathcal{L}\left[u;\mathsf{A},\mathsf{B}\right]=\nabla \mathcal{P}$ and discriminant Δ(h, g),” Nonlinear Anal., vol. 173, pp. 209–219, 2018, https://doi.org/10.1016/j.na.2018.04.002.Search in Google Scholar

[21] L. Ni, “The entropy formula for linear heat equation,” J. Geom. Anal., vol. 14, no. 1, pp. 85–98, 2004. https://doi.org/10.1007/bf02921867.Search in Google Scholar

[22] R. Schoen and S. T. Yau, Lectures on Differential Geometry, International Press, 1994.Search in Google Scholar

[23] P. Souplet and Q. S. Zhang, “Sharp gradient estimate and Yau’s Liouville theorem for the heat equation on noncompact manifolds,” Bull. Lond. Math. Soc., vol. 38, pp. 1045–1053, 2006, https://doi.org/10.1112/s0024609306018947.Search in Google Scholar

[24] K. T. Sturm, “Super-Ricci flows for metric measure spaces,” J. Funct. Anal., vol. 275, pp. 3504–3569, 2018, https://doi.org/10.1016/j.jfa.2018.07.014.Search in Google Scholar

[25] A. Taheri, Function Spaces and Partial Differential Equations, Vol. I, Oxford Lecture Series in Mathematics and its Applications, vol. 40, OUP, 2015.Search in Google Scholar

[26] A. Taheri, Function Spaces and Partial Differential Equations, Vol. II, Oxford Lecture Series in Mathematics and its Applications, vol. 41, OUP, 2015.10.1093/acprof:oso/9780198733157.001.0001Search in Google Scholar

[27] A. Taheri, “Liouville theorems and elliptic gradient estimates for a nonlinear parabolic equation involving the Witten Laplacian,” Adv. Calc. Var., vol. 16, pp. 425–441, 2023, https://doi.org/10.1515/acv-2020-0099.Search in Google Scholar

[28] A. Taheri, “Gradient estimates for a weighted Γ-nonlinear parabolic equation coupled with a super Perelman-Ricci flow and implications,” Potential Anal., vol. 59, pp. 311–335, 2023, https://doi.org/10.1007/s11118-021-09969-2.Search in Google Scholar

[29] A. Taheri and V. Vahidifar, “On multiple solutions to a family of nonlinear elliptic systems in divergence form combined with an incompressibility constraint,” Nonlinear Anal., vol. 221, 2022, Art. no. 112889, https://doi.org/10.1016/j.na.2022.112889.Search in Google Scholar

[30] A. Taheri and V. Vahidifar, “Gradient estimates for a nonlinear parabolic equation on smooth metric measure spaces with evolving metrics and potentials,” Nonlinear Anal., vol. 232, 2023, Art. no. 113255, https://doi.org/10.1016/j.na.2023.113255.Search in Google Scholar

[31] A. Taheri and V. Vahidifar, “Souplet-Zhang and Hamilton type gradient estimates for nonlinear elliptic equations on smooth metric measure spaces,” Mathematika, vol. 69, pp. 751–779, 2023, https://doi.org/10.1112/mtk.12208.Search in Google Scholar

[32] A. Taheri and V. Vahidifar, “Differential Harnack estimates for a weighted nonlinear parabolic equation under a super Perelman-Ricci flow and implications,” Proc. Roy. Soc. Edinb. A, 2024.10.1007/s42985-023-00269-5Search in Google Scholar

[33] W. Wang, “Harnack inequality, heat kernel bounds and eigenvalue estimate under integral Ricci curvature bounds,” J. Differ. Equ., vol. 269, pp. 1243–1277, 2020, https://doi.org/10.1016/j.jde.2020.01.003.Search in Google Scholar

[34] J. Y. Wu, “Elliptic gradient estimates for a weighted heat equation and applications,” Math. Z., vol. 280, pp. 451–468, 2015, https://doi.org/10.1007/s00209-015-1432-9.Search in Google Scholar

[35] J. Y. Wu, “Gradient estimates for nonlinear parabolic equation and Liouville theorems,” Manuscripta Math., vol. 159, pp. 511–547, 2019, https://doi.org/10.1007/s00229-018-1073-5.Search in Google Scholar

[36] Q. S. Zhang, Sobolev Inequalities, Heat Kernels Under Ricci Flow and the Poincaré Conjecture, CRC Press, 2011.10.1201/EBK1439834596Search in Google Scholar

[37] J. Li, “Gradient estimates and Harnack inequalities for nonlinear parabolic and nonlinear elliptic equations on Riemannian manifolds,” J. Funct. Anal., vol. 100, pp. 233–256, 1991, https://doi.org/10.1016/0022-1236(91)90110-q.Search in Google Scholar

[38] D. Bakry and M. Émery, “Diffusions hypercontractives,” in Séminaire de Probabilités XIX 1983/84. Lecture Notes in Mathematics, vol. 1123, J. Azéma and M Yor, Eds., Berlin, Heidelberg, Springer.Search in Google Scholar

[39] D. Bakry, I. Gentil, and M. Ledoux, Analysis and Geometry of Markov Diffusion Operators, A Series of Comprehensive Studies in Mathematics, vol. 348, Springer, 2012.Search in Google Scholar

[40] L. Gross, “Logarithmic Sobolev inequalities,” Am. J. Math., vol. 97, pp. 1061–1083, 1976, https://doi.org/10.2307/2373688.Search in Google Scholar

[41] C. Villani, Optimal Transport: Old and New, A Series of Comprehensive Studies in Mathematics, vol. 338, Springer, 2008.Search in Google Scholar

[42] F. Y. Wang, Analysis for Diffusion Processes on Riemannian Manifolds, Advanced Series on Statistical Science & Probability, vol. 18, World Scientific, 2013.10.1142/8737Search in Google Scholar

[43] J. Lott, “Some geometric properties of the Bakry-Émery Ricci tensor,” Comment Math. Helv., vol. 78, pp. 865–883, 2003, https://doi.org/10.1007/s00014-003-0775-8.Search in Google Scholar

[44] H. D. Cao, B. Chow, S. C. Chu, and S. T. Yau, Collected Papers on Ricci Flow, Surveys in Geometry and Topology, vol. 37, International Press, 2003.Search in Google Scholar

[45] R. Hamilton, “The formation of singularities in the Ricci flow,” Surv. Differ. Geom., vol. 2, pp. 7–136, 1995, https://doi.org/10.4310/sdg.1993.v2.n1.a2.Search in Google Scholar

[46] J. M. Lee and T. H. Parker, “The Yamabe problem,” Bull. Am. Math. Soc., vol. 17, pp. 37–91, 1987, https://doi.org/10.1090/s0273-0979-1987-15514-5.Search in Google Scholar

[47] P. Mastrolia, M. Rigoli, and A. G. Setti, Yamabe Type Equations on Complete Non-Compact Manifolds, Basel, Springer, 2012.10.1007/978-3-0348-0376-2Search in Google Scholar

[48] J. Case, “A Yamabe-type problem on smooth metric measure spaces,” J. Differ. Geom., vol. 101, pp. 467–505, 2015, https://doi.org/10.4310/jdg/1445518921.Search in Google Scholar

[49] N. T. Dung, N. N. Khanh, and Q. A. Ngô, “Gradient estimates for f-heat equations driven by Lichnerowicz’s equation on complete smooth metric measure spaces,” Manuscripta Math., vol. 155, pp. 471–501, 2018, https://doi.org/10.1007/s00229-017-0946-3.Search in Google Scholar

[50] S. M. Allen and J. W. Cahn, “A microscopic theory for antiphase boundary motion and and its applications to antiphase domain coarsening,” Acta Metall., vol. 27, pp. 1085–1095, 1979, https://doi.org/10.1016/0001-6160(79)90196-2.Search in Google Scholar

[51] R. A. Fisher, “The wave of advance of advantageous genes,” Ann Eugen., vol. 7, pp. 355–369, 1937, https://doi.org/10.1111/j.1469-1809.1937.tb02153.x.Search in Google Scholar

[52] A. N. Kolmogorov, I. G. Petrovskii, and N. S. Piskunov, “Etude de l’équation de la diffusion avec croissance de la quantité de matière et son application a un probléme biologique,” Bull. Univ. d’Etatá, Moscou, Ser. Int. A, vol. 1, pp. 1–26, 1937.Search in Google Scholar

[53] M. Bǎileşteanu, “A Harnack inequality for the parabolic Allen-Cahn equation,” Ann. Glob. Anal. Geom., vol. 51, pp. 367–378, 2017, https://doi.org/10.1007/s10455-016-9540-2.Search in Google Scholar

[54] X. D. Cao, B. W. Liu, I. Pendleton, and A. Ward, “Differential Harnack estimates for Fisher’s equation,” Pac. J. Math., vol. 290, pp. 273–300, 2017, https://doi.org/10.2140/pjm.2017.290.273.Search in Google Scholar

[55] L. A. Cafarreli, B. Gidas, and J. Spruck, “Asymptotic symmetry and local behavior of semilinear elliptic equations with critical Sobolev growth,” Commun. Pure Appl. Math., vol. 42, pp. 271–297, 1989, https://doi.org/10.1002/cpa.3160420304.Search in Google Scholar

[56] M. Ghergu, S. Kim, and H. Shahgholian, “Exact behaviour around isolated singularity for semilinear elliptic equations with a log-type nonlinearity,” Adv. Nonlinear Anal., vol. 8, pp. 995–1003, 2019, https://doi.org/10.1515/anona-2017-0261.Search in Google Scholar

[57] Y. Choquet-Bruhat, General Relativity and the Einstein Equations, Oxford Mathematical Monographs, OUP, 2009.10.1093/acprof:oso/9780199230723.001.0001Search in Google Scholar

[58] L. Ma, “Gradient estimates for a simple elliptic equation on complete noncompact Riemannian manifolds,” J. Funct. Anal., vol. 241, pp. 374–382, 2006, https://doi.org/10.1016/j.jfa.2006.06.006.Search in Google Scholar

[59] X. F. Song and L. Zhao, “Gradient estimates for the elliptic and parabolic Lichnerowicz equations on compact manifolds,” Z. Angew. Math. Phys., vol. 61, pp. 655–662, 2010, https://doi.org/10.1007/s00033-009-0047-6.Search in Google Scholar

[60] R. Müller, Differential Harnack Inequalities and the Ricci Flow, EMS Series of Lectures in Mathematics, EMS, 2006.10.4171/030Search in Google Scholar

[61] G. Perelman, “The entropy formula for the Ricci Flow and its geometric application,” arXiv: math. DG/0211159v1, 2002.Search in Google Scholar

[62] D. Bakry, “L’hypercontractivité et son utilisation en théorie des semigroupes,” in Lecture Notes in Math., vol. 1581, Berlin, New York, Springer-Verlag, 1994, pp. 1–114.10.1007/BFb0073872Search in Google Scholar

[63] L. F. Wang, “The upper bound of the Lμ2${L}_{\mu }^{2}$-spectrum,” Ann. Global Anal. Geom., vol. 37, pp. 393–402, 2010. https://doi.org/10.1007/s10455-009-9193-5.Search in Google Scholar

[64] L. F. Wang, “A splitting theorem for the weighted measure,” Ann. Global Anal. Geom., vol. 42, pp. 79–89, 2012, https://doi.org/10.1007/s10455-011-9302-0.Search in Google Scholar

[65] K. Brighton, “A Liouville-type theorem for smooth metric measure spaces,” J. Geom. Anal., vol. 23, pp. 562–570, 2013, https://doi.org/10.1007/s12220-011-9253-5.Search in Google Scholar

[66] G. Wei and W. Wylie, “Comparison geometry for the Bakry-Émery Ricci tensor,” J. Differ. Geom., vol. 83, pp. 377–405, 2009, https://doi.org/10.4310/jdg/1261495336.Search in Google Scholar

Received: 2022-12-16
Accepted: 2023-12-18
Published Online: 2024-04-11

© 2024 the author(s), published by De Gruyter, Berlin/Boston

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

Downloaded on 15.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2023-0120/html
Scroll to top button