1 Introduction

Let \((M^m,g)\) be a compact Riemannian manifolds with nonempty boundary \(\partial M\) and \((N^n, {\tilde{g}})\) a complete Riemannian manifold without boundary. Let \(d: N \times N \rightarrow {\mathbb {R}} \) be the distance function on N and \(B_{(1+\sigma )R}(p):= \{ q \in N: d(p, q) \le (1+\sigma )R \}\) a regular ball in N, that is, disjoint from the cut locus of its center p and of radius \((1+\sigma )R < \frac{\pi }{2\sqrt{\kappa }}\), where \(\kappa = \max \{ 0, \sup _{B_{(1+\sigma )R}(p)}K_N \}\) and \(\sup _{B_{(1+\sigma )R}(p)}K_N\) is an upper bound of the sectional curvature K of N on \(B_{(1+\sigma )R}(p)\), and \(\sigma >0\) is any given constant.

Let \(V \in \Gamma (TM)\), \(T \in \Gamma (\otimes ^{1, 2}TN)\). We call a map \(u: M \rightarrow N \) a VT-harmonic map if u satisfies

$$\begin{aligned} \tau (u)+du(V)+ \mathrm{Tr}_g T(du, du) = 0, \end{aligned}$$
(1.1)

where \(\tau (u)=\text{ tr }D du\) is the tension field of the map u. This is a generalization the notion of a V-harmonic map that has been studied in recent years as a common framework including Hermitian, affine and Weyl harmonic maps into Riemannian manifolds, that is, the domain possessed a connection different from the Levi-Civita connection, but the target was a Riemannian manifold with its Levi-Civita connection. This generalized the standard harmonic map system \(\tau (u)=0\) to a system of the form \(\tau (u)+du(V)=0\) with a vector field V on the domain. Here, we want to consider targets that are of the same type as the domain. That leads to the system (1.1) with an additional term arising from a 2-tensor T on the target. As this new term \(\mathrm{Tr}_g T(du, du)\), in contrast to the term du(V), is analytically of the same weight as the elliptic operator \(\tau (u)\) (which includes the Laplace–Beltrami operator of the domain), this makes the analysis more difficult and subtle. This is the problem that we are addressing in this paper.

In local coordinates \(\{ x^\alpha \}\) on M and \(\{ y^i \}\) on N, respectively, we can write (1.1) as

$$\begin{aligned} \Delta _M u^i + \Gamma ^{i}_{jk}(u)\frac{\partial u^j}{\partial x^\alpha }\frac{\partial u^k}{\partial x^\beta }g^{\alpha \beta } + V^\alpha \frac{\partial u^i}{\partial x^\alpha } + T^{i}_{jk}(u)\frac{\partial u^j}{\partial x^\alpha }\frac{\partial u^k}{\partial x^\beta }g^{\alpha \beta }=0, \end{aligned}$$
(1.2)

where \(\Delta _M\) is the Laplacian on (Mg), \(\Gamma ^{i}_{jk}\) stands for the Christoffel symbols of \((N, {\tilde{g}})\), \(V:=V^\alpha \frac{\partial }{\partial x^\alpha }\) and \(T:=T^{i}_{jk}\frac{\partial }{\partial y^i}\otimes dy^j\otimes dy^k\). This is a second-order semilinear elliptic system on the manifold (Mg).

As is already the case for V-harmonic maps, in general, (1.1) is neither in divergence form, nor has a variational structure. Chen et al. [5] established a Jäger–Kaul type maximum principle for V-harmonic maps by using the method of [8], and combining this with the continuity method, the existence of V-harmonic maps into a regular ball could be proved. Therefore, it is natural to ask whether a maximum principle holds for VT-harmonic maps. However, the case of VT-harmonic maps is harder to deal with than V-harmonic maps since we now have an additional quadratic term arising from the tensor T. Due to this additional structure, the construction of the elliptic operator in [5] is no longer valid in our case. To overcome this difficulty, we use another construction as in [7] to compensate this term and obtain the following maximum principle for VT-harmonic maps:

Theorem 1

Let \(u_1, u_2 \in C^0(M, N)\) be two VT-harmonic maps into a geodesic ball \(B_{R}(p)\). For appropriate \(\sigma \) and R, there exists a constant \(C_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla T| + \max |T| \le C_0, \end{aligned}$$
(1.3)

then the function \(\Theta : M \rightarrow {\mathbb {R}}\) defined by

$$\begin{aligned} \Theta := \frac{q_{\frac{\kappa }{4}}(\rho )}{(q_\kappa ((1+\sigma )R)-q_\kappa (\rho _1))^{\frac{1}{2}}\cdot (q_\kappa ((1+\sigma )R)-q_\kappa (\rho _2))^{\frac{1}{2}}} \end{aligned}$$
(1.4)

satisfies the maximum principle, namely

$$\begin{aligned} \max _M \Theta \le \max _{\partial M}\Theta . \end{aligned}$$

Here the expression of \(q_\kappa \) is given in Sect. 2, and \(\rho := d(u_1, u_2), \rho _i:= d(p, u_i), i=1,2.\)

In particular, if \(u_1=u_2\) on the boundary \(\partial M\), then \(u_1\equiv u_2\) on M.

Remark

The explicit expression of the constant \(C_0\) in the above and in the subsequent results can be seen in (3.5). Importantly, \(C_0\rightarrow \infty \) for \(R\rightarrow 0\). Thus, we can also satisfy the condition on T by making the target ball sufficiently small.

For the heat flow of VT-harmonic maps, an analogous result holds. For \(T>0\), we set

$$\begin{aligned} M_T:= M\times [0, T] \end{aligned}$$

and denote the parabolic boundary of \(M_T\) by

$$\begin{aligned} \partial _p M_T:=(M\times \{0\})\cup (\partial M\times [0, T]). \end{aligned}$$

We consider the heat flow of VT-harmonic maps

$$\begin{aligned} \partial _t u= \tau (u)+du(V) + \mathrm{Tr}_g T(du, du) \end{aligned}$$
(1.5)

and have

Theorem 2

Let \(u_1, u_2 \in C^0(M, N)\) be two solutions of heat flow Eq. (1.5) for VT-harmonic maps into a geodesic ball \(B_{R}(p)\). For appropriate \(\sigma \) and R, there exists a constant \(C_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla T| + \max |T| \le C_0, \end{aligned}$$

then the function \(\Theta : M_T \rightarrow {\mathbb {R}}\) defined by (1.4) with M replaced by \(M_T\) satisfies the maximum principle:

$$\begin{aligned} \max _{M_T} \Theta \le \max _{\partial _p M_T}\Theta . \end{aligned}$$

In particular, if \(u_1=u_2\) on the boundary \(\partial _p M_T\), then \(u_1\equiv u_2\) on \(M_T\).

As an application of the above maximum principle, we obtain the existence of VT-harmonic maps into a geodesic ball.

Theorem 3

Let \(M, N, V, T, B_R(p)\) be as in Theorem 2. Suppose \(u_0 \in H^{2, q}(M, N)(q>m)\) with \(u_0(M) \subset B_R(p)\). For appropriate \(\sigma \) and R, there exists a constant \(C_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla T| + \max |T| \le C_0, \end{aligned}$$

then the initial boundary value problem

$$\begin{aligned} \left\{ \begin{array}{rll} \displaystyle &{}\partial _tu= \tau (u) + du(V) + \mathrm{Tr}_g T(du, du), \\ \displaystyle &{} u-u_0 \in H^{2, q}_0(M, N),\quad u(0)=u_0, \quad u(M\times [0,\infty ))\subset B_R(p), \end{array} \right. \end{aligned}$$
(1.6)

admits a unique global solution u which subconverges to a unique solution \(u \in H^{2, q}(M, N)\) of the Dirichlet problem

$$\begin{aligned} \left\{ \begin{array}{rll} \displaystyle &{} \tau (u) + du(V) + \mathrm{Tr}_g T(du, du) = 0, \\ \displaystyle &{} u-u_0 \in H^{2, q}_0(M, N), \end{array} \right. \end{aligned}$$
(1.7)

such that \(u(M)\subset B_R(p)\).

Furthermore, based on Theorem 3, we shall also establish the existence of VT-harmonic maps from complete noncompact Riemannian manifolds by using a gradient estimate and the compact exhaustion method.

Theorem 4

Let \((M^m, g)\) be a complete noncompact Riemannian manifold and \((N^n, {\tilde{g}})\) be a complete Riemannian manifold with sectional curvature bounded above by a positive constant \(\kappa \). Let \(B_R(p)\) be a geodesic ball with radius \(R< \frac{\pi }{2(1+\sigma )\sqrt{\kappa }}\) and \(u_0: M\rightarrow N\) a smooth map with \(u_0(M) \subset B_R(p)\). Suppose \(\Vert V\Vert _{L^\infty }< +\infty .\)

For appropriate \(\sigma \) and R, there exists a constant \(C'_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla T| + \max |T| \le C'_0, \end{aligned}$$

then there exists a VT-harmonic map \(u \in C^{\infty }(M, N)\) homotopic to \(u_0\) such that \(u(M)\subset B_R(p)\).

2 Preliminaries

Let us first give some notations:

$$\begin{aligned} s_\kappa (t):= & {} \left\{ \begin{array}{ll} \displaystyle t &{} \quad \kappa =0 \\ \displaystyle \frac{1}{\sqrt{\kappa }}\sin \sqrt{\kappa }t&{} \quad \kappa>0, \end{array} \right. \qquad q_\kappa (t):=\left\{ \begin{array}{ll} \displaystyle \frac{t^2}{2} &{} \quad \kappa =0 \\ \displaystyle \frac{1}{\kappa }(1-\cos \sqrt{\kappa }t)&{} \quad \kappa>0. \end{array} \right. \\ a_\kappa (t):= & {} \left\{ \begin{array}{ll} \displaystyle 0 &{} \quad \mathrm{if} \ \ t=0\\ \displaystyle \frac{1-s'_\kappa (t)}{s_\kappa (t)} &{} \quad \mathrm{if} \,\, t>0, \end{array} \right. \quad \quad b_\kappa (t):=\left\{ \begin{array}{ll} \displaystyle 0 &{}\quad \mathrm{if} \ \ t=0\\ \displaystyle \frac{1-s'_\kappa (t)}{2s_\kappa (t)}\left( 1+ \frac{t}{s_\kappa (t)} \right) \ &{} \quad \mathrm{if} \ \ t>0. \end{array}\right. \end{aligned}$$

In local coordinates \(\{ x^\alpha \}\) on M and \(\{ y^i \}\) on N, respectively, the energy density of u is

$$\begin{aligned} e(u):= g^{\alpha \beta }{\tilde{g}}_{ij}(u(x))\frac{\partial u^i}{\partial x^\alpha }\frac{\partial u^j}{\partial x^\beta }. \end{aligned}$$

Assume the metric of N satisfies:

$$\begin{aligned} 0<{\widetilde{\lambda }}(y)(\delta _{ij})\le ({\widetilde{g}}_{ij}(y))\le {\widetilde{\Lambda }}(y)(\delta _{ij}), \quad \forall y\in N. \end{aligned}$$

Denote \(\lambda :=\min \nolimits _N{{\widetilde{\lambda }}}\) and \(\Lambda :=\max \nolimits _N{{\widetilde{\Lambda }}}\)

\(\forall y_1, y_2 \in B_R(p)\), there exists a unit speed geodesic \(\gamma : [0, \rho ] \rightarrow B_R(p)\subset N\) with \(\gamma (0)= y_1, \gamma (\rho )=y_2\), where \(\rho =\mathrm{dist}(y_1, y_2)\). For any \(v_j \in T_{y_j}N, j=1, 2\), let X be the unique Jacobi field along \(\gamma \) with \(X(0)=v_1, X(\rho )= v_2\). Then, we define a pseudo-distance

$$\begin{aligned} \delta (v_1, v_2):=\left\{ \begin{array}{ll} \displaystyle \left( \rho \int ^{\rho }_{0}|{\dot{X}}|^2 \right) ^{\frac{1}{2}}&{}\quad \mathrm{if} \ \ \rho >0,\\ \displaystyle |v_1 - v_2| &{}\quad \mathrm{if} \ \ \rho =0. \end{array} \right. \end{aligned}$$

Another pseudo-distance is given by

$$\begin{aligned} \delta _0(v_1, v_2) := \left| v_1 - \bar{{\bar{v}}}_2 \right| , \end{aligned}$$

where \(\bar{{\bar{v}}}_2 \in T_{y_1}N\) stands for the vector obtained by parallel displacement of \(v_2 \in T_{y_2}N\) along \(\gamma \). Let \(L(T_x M, T_y N)\) be the space of all linear maps from \(T_x M\) to \(T_y N\). The pseudo-distance \(\delta \) on the tangent bundle can be extended to a pseudo-distance on the fibers, that is, for \(q_1, q_2 \in \cup _{y\in B_R(p)}L(T_xM, T_yN) \) (disjoint union), we define their pseudo-distance as

$$\begin{aligned} \delta (q_1, q_2) := \left( \sum _{\alpha =1}^{m}\delta ^2(q_1(e_\alpha ), q_2(e_\alpha ))\right) ^{\frac{1}{2}}, \end{aligned}$$

where \(\{ e_1, \ldots , e_m \}\) is an orthonormal base for \(T_{x}M\).

We have the following relationship between these two pseudo-distances:

Lemma 1

([4]) There is a positive constant C depending only on \(B_R(p)\) and the geometry of N such that for any \(y_j \in B_R(p)\) and \(v_j \in T_{y_j}N, j=1, 2\), we have

$$\begin{aligned} \delta _{0}^2(v_1, v_2) - C(|v_1|^2+|v_2|^2)\rho ^2 \le \delta ^2(v_1, v_2)\le \delta _{0}^2(v_1, v_2) + C(|v_1|^2+|v_2|^2)\rho ^2. \end{aligned}$$

Remark 1

In fact, by the proof in [4] and using a well-known expression of the curvature operator (see, e.g., Lemma 4.3.3 in [12]), it is not hard to see that if the sectional curvature K on \(B_R(p)\) satisfies \( \theta \le \left. K\right| _{B_R(p)} \le \kappa \) for a constant \(\theta < 0\), then the constant C can be expressed as \(14(\kappa - \theta )\).

The following estimates will also be important for us:

Lemma 2

([7]) Let (Mg) be a compact Riemannian manifolds with nonempty boundary \(\partial M\) and \((N, {\tilde{g}})\) a complete manifold without boundary and \(B_R(p)\) a regular ball in N. Let

$$\begin{aligned} \begin{aligned} g_1:=&q_\kappa \circ d(p, \cdot ): B_R(p) \rightarrow {\mathbb {R}},\\ h:=&q_{\frac{\kappa }{4}}\circ d: B_R(p) \times B_R(p) \rightarrow {\mathbb {R}}. \end{aligned} \end{aligned}$$

Then,

$$\begin{aligned} \nabla ^2g_1(u,u) \ge s'_\kappa (\tau )|u|^2 \end{aligned}$$
(2.1)

hold for \(u \in T_x N, x \in B_R(p) \) and \(\tau := d(p, x)\).

$$\begin{aligned} \nabla ^2h(v, v) \ge -s_{\frac{\kappa }{4}}(\rho )a_\kappa (\rho )\sum _{i=1}^2|v_i|^2 \end{aligned}$$
(2.2)

and

$$\begin{aligned} \nabla ^2h(v,v)\ge s'_{\frac{\kappa }{4}}(\rho )\delta ^2(v_1, v_2) - s_{\frac{\kappa }{4}}(\rho )b_\kappa (\rho )\sum _{i=1}^2|v_i|^2 \end{aligned}$$
(2.3)

holds for \(v=v_1\oplus v_2, v_j \in T_{y_j}N, y_j \in B_R(p), j=1, 2, \rho =\mathrm{dist}(y_1, y_2)\).

3 The maximum principle

Proof of Theorem 1

Let

$$\begin{aligned} \begin{aligned} \psi (x):=&q_{\frac{\kappa }{4}}\circ d(u_1(x), u_2(x)),\\ \psi _i(x):=&q_\kappa \circ d(p, u_i(x)), \ \ \ i=1, 2, \\ \varPhi (x):=&\frac{1}{2}\sum _{i=1}^2 \omega \circ \psi _i(x), \ \ \ \mathrm{where} \ \ \ \omega (t):=-\log (q_\kappa ((1+\sigma )R)-t). \end{aligned} \end{aligned}$$

We consider the operator

$$\begin{aligned} {\mathcal {L}}_V(\cdot ):= e^\varPhi \cdot \mathrm{div}(e^{-2\varPhi }\nabla \cdot )+e^{-\varPhi }\cdot V(\cdot ). \end{aligned}$$

By direct computation, we obtain

$$\begin{aligned} \begin{aligned} {\mathcal {L}}_V (\Theta ) =&{\mathcal {L}}_V (e^\varPhi \cdot \psi )=\Delta \psi + \psi (\Delta \varPhi - |\nabla \varPhi |^2) + \psi V(\varPhi ) + V(\psi )\\ =&\Delta _V \psi + \psi (\Delta _V \varPhi - |\nabla \varPhi |^2). \end{aligned} \end{aligned}$$

Define \(U, U_1, U_2: M \rightarrow N\times N\) by

$$\begin{aligned} U(x):= (u_1(x), u_2(x)), \ \ \ U_i(x):=(p, u_i(x)), \ \ \ i=1, 2. \end{aligned}$$

Let \(\nu :=\frac{\kappa }{4}, h:= q_\nu \circ d, \phi :=q_\kappa \circ d,\) then \(\psi =h\circ U, \psi _i=\phi \circ U_i\).

For any \(x \in M\), we let \({\tilde{\gamma }}\) be the unique geodesic connecting \(u_1(x)\) and \(u_2(x)\). Choosing a parallel orthonormal frame \(\{ E_i(t) \}\) along \({\tilde{\gamma }}\) with \(E_1 = {\tilde{\gamma }}'\), and a local orthonormal frame \(\left. \{ e_\alpha \}\right| _{\alpha =1}^{m}\) around x, assuming that \(\frac{\partial }{\partial y^i} := a^{j}_iE_j\), we have

$$\begin{aligned} \begin{aligned}&\delta _0(T(du_1(e_\alpha ), du_1(e_\alpha )), T(du_2(e_\alpha ), du_2(e_\alpha )))\\&\quad =\delta _0\left( (u_1)^{i}_{\alpha }(u_1)^{j}_{\alpha }T^{r}_{ij}(u_1)\frac{\partial }{\partial y^r}(u_1), (u_2)^{i}_{\alpha }(u_2)^{j}_{\alpha }T^{r}_{ij}(u_2)\frac{\partial }{\partial y^r}(u_2)\right) \\&\quad = \delta _0\left( (u_1)^{i}_{\alpha }(u_1)^{j}_{\alpha }T^{r}_{ij}(u_1)a^{\mu }_{r}(u_1)E_\mu (u_1), (u_2)^{i}_{\alpha }(u_2)^{j}_{\alpha }T^{r}_{ij}(u_2)a^{\mu }_{r}(u_2)E_\mu (u_2)\right) \\&\quad \le \sum _{\mu } \left| T^{r}_{ij}(u_1)a^{\mu }_r(u_1)(u_1)^{i}_{\alpha }(u_1)^{j}_\alpha - T^{r}_{ij}(u_2)a^{\mu }_r(u_2)(u_2)^{i}_{\alpha }(u_2)^{j}_\alpha \right| \\&\quad \le \sum _{\mu } \left| (T^{r}_{ij}(u_1)-T^{r}_{ij}(u_2))a^{\mu }_r (u_1)(u_1)^{i}_{\alpha }(u_1)^{j}_\alpha \right| \\&\qquad + \sum _{\mu } \left| T^{r}_{ij}(u_2)( a^{\mu }_r (u_1) - a^{\mu }_r (u_2) )(u_1)^{i}_{\alpha }(u_1)^{j}_\alpha \right| \\&\qquad + \sum _{\mu }\left| T^{r}_{ij}(u_2)a^{\mu }_r(u_2) (u_1)^{j}_\alpha \left( (u_1)^{i}_\alpha - (u_2)^{i}_\alpha \right) \right| \\&\qquad + \sum _{\mu }\left| T^{r}_{ij}(u_2)a^{\mu }_r(u_2) (u_2)^{i}_{\alpha } \left( (u_1)^{j}_\alpha - (u_2)^{j}_\alpha \right) \right| . \end{aligned} \end{aligned}$$

Denote \(A= (a^{k}_i)\), then \(AA^T=G:=({\tilde{g}}_{ik})\). Since

$$\begin{aligned} \sum _{i,k}|a^{k}_i|^2 = \Vert A\Vert ^2 = \mathrm{tr}(AA^T) =\mathrm{tr}(G)\le n\Lambda , \quad \quad \left| (u_1)^{i}_{\alpha }(u_1)^{j}_\alpha \right| \le \frac{1}{\lambda }|du_1|^2, \end{aligned}$$

we then obtain

$$\begin{aligned} \begin{aligned}&\delta _0(T(du_1(e_\alpha ), du_1(e_\alpha )), T(du_2(e_\alpha ), du_2(e_\alpha )))\\&\quad \le \sum _{\mu =1}^n\left( \frac{\sqrt{n\Lambda }}{\lambda } \rho \max | \nabla T| |du_1|^2 + \frac{{\widetilde{C}}_1}{\lambda }\rho \max |T||du_1|^2 \right. \\&\left. \qquad + \frac{\sqrt{n\Lambda }}{\lambda }\max |T|(|du_1|+|du_2|) |du_1-du_2|\right) \\&\quad = \frac{n\sqrt{n\Lambda }}{\lambda } \rho \max | \nabla T| |du_1|^2 + \frac{n{\widetilde{C}}_1}{\lambda }\rho \max |T||du_1|^2 \\&\qquad + \frac{n\sqrt{n\Lambda }}{\lambda }\max |T|(|du_1|+|du_2|) |du_1-du_2|, \end{aligned} \end{aligned}$$

where \({\widetilde{C}}_1>0\) is a constant depending only on the bound of \((da_{r}^\mu )\) on \(B_{\frac{\pi }{2\sqrt{\kappa }}}(p)\). By Lemma 1, we get

$$\begin{aligned} \begin{aligned}&\delta (T(du_1(e_\alpha ), du_1(e_\alpha )), T(du_2(e_\alpha ), du_2(e_\alpha )))\\&\quad \le \delta _0(T(du_1(e_\alpha ), du_1(e_\alpha )), T(du_2(e_\alpha ), du_2(e_\alpha ))) + \sqrt{C}\rho (\left| T(du_1(e_\alpha ), du_1(e_\alpha )) \right| \\&\qquad + \left| T(du_2(e_\alpha ), du_2(e_\alpha )) \right| ) \\&\quad \le \frac{n\sqrt{n\Lambda }}{\lambda } \rho \max | \nabla T| \sum _{i=1}^2 |du_i|^2 + \left( \frac{n{\widetilde{C}}_1}{\lambda }+\sqrt{C}\right) \rho \max |T|\sum _{i=1}^2 |du_i|^2\\&\qquad + \frac{n\sqrt{n\Lambda }}{\lambda }\max |T|(|du_1|+|du_2|) |du_1-du_2|, \end{aligned} \end{aligned}$$

where \(C=14(\kappa -\theta )\), and the constant \(\theta \) is a lower bound of the sectional curvature of N on \(B_{\frac{\pi }{2\sqrt{\kappa }}}(p)\). The Cauchy inequality implies that

$$\begin{aligned}&\frac{n\sqrt{n\Lambda }}{\lambda }\max |T||du_i| |du_1-du_2|s_\nu (\rho )\le \frac{\varepsilon _1n^2}{2}|du_1-du_2|^2 \\&\quad + \frac{n\Lambda }{2\varepsilon _1 \lambda ^2}s_\nu ^2(\rho )\max |T|^2 |du_i|^2. \end{aligned}$$

By using the formula (2.13) in [7], it follows that

$$\begin{aligned} \begin{aligned}&\langle (\nabla h)\circ U, -T(dU(e_\alpha ), dU(e_\alpha ))\rangle \\&\quad = -s_\nu (\rho ) \langle (\nabla d)\circ U, T(dU(e_\alpha ), dU(e_\alpha )) \rangle \\&\quad = -s_\nu (\rho ) \langle {\tilde{e}}_1(U)\oplus {\tilde{e}}_2(U), T(du_1(e_\alpha ), du_1(e_\alpha )) \oplus T(du_2(e_\alpha ), du_2(e_\alpha )) \rangle \\&\quad \ge -s_\nu (\rho ) \delta (T(du_1(e_\alpha ), du_1(e_\alpha )), T(du_2(e_\alpha ), du_2(e_\alpha )))\\&\quad \ge -s_\nu (\rho ) \left\{ \frac{n\sqrt{n\Lambda }}{\lambda } \rho \max | \nabla T| \sum _{i=1}^2 |du_i|^2 + \left( \frac{n{\widetilde{C}}_1}{\lambda }+\sqrt{C}\right) \rho \max |T|\sum _{i=1}^2 |du_i|^2 \right\} \\&\qquad -\varepsilon _1 n^2|du_1-du_2|^2 - \frac{n\Lambda }{2\varepsilon _1\lambda ^2}s_\nu ^2(\rho )\max |T|^2 \sum _{i=1}^2|du_i|^2, \end{aligned} \end{aligned}$$

where \({\tilde{e}}_1 = -{\tilde{\gamma }}'(0), {\tilde{e}}_2 = {\tilde{\gamma }}'(\rho )\). Since

$$\begin{aligned} \begin{aligned}&\delta ^{2}_{0}(du_1(e_\alpha ), du_2(e_\alpha )) = \delta ^{2}_0 \left( (u_1)^{i}_\alpha \frac{\partial }{\partial y^i}(u_1), (u_2)^{i}_\alpha \frac{\partial }{\partial y^i}(u_2)\right) \\&\quad = \delta ^{2}_0\left( (u_1)^{i}_\alpha a^{j}_i(u_1)E_j(u_1), (u_2)^{i}_\alpha a^{j}_i(u_2)E_j(u_2) \right) = \sum _j \left[ (u_1)^{i}_\alpha a^{j}_i(u_1) - (u_2)^{i}_\alpha a^{j}_i(u_2) \right] ^2 \\&\quad = \sum _j \left[ (u_1)^{i}_\alpha (a^{j}_i(u_1)-a^{j}_i(u_2)) + a^{j}_i(u_2)((u_1)^{i}_\alpha - (u_2)^{i}_\alpha ) \right] ^2 \\&\quad \ge \frac{1}{2}\sum _j \left[ a^{j}_i(u_2)((u_1)^{i}_\alpha - (u_2)^{i}_\alpha ) \right] ^2 - \sum _j \left[ (u_1)^{i}_\alpha (a^{j}_i(u_1)-a^{j}_i(u_2)) \right] ^2 \\&\quad = \frac{1}{2}\sum _j\sum _{ik}{\tilde{g}}_{ik}(u_2)((u_1)^{i}_\alpha - (u_2)^{i}_\alpha )((u_1)^{k}_\alpha - (u_2)^{k}_\alpha ) - \sum _j \left[ (u_1)^{i}_\alpha (a^{j}_i(u_1)-a^{j}_i(u_2)) \right] ^2 \\&\quad \ge \frac{n\lambda }{2} | (du_1-du_2)(e_\alpha ) |^2 - n{\widetilde{C}}_1 \rho ^2 |du_1(e_\alpha )|^2, \end{aligned} \end{aligned}$$

then by Lemma 1, we have

$$\begin{aligned} \begin{aligned}&\delta ^{2}(du_1(e_\alpha ), du_2(e_\alpha )) \ge \delta ^{2}_{0}(du_1(e_\alpha ), du_2(e_\alpha )) -C\left( |du_1(e_\alpha )|^2+|du_2(e_\alpha )|^2 \right) \rho ^2 \\&\quad \ge \frac{n\lambda }{2} | (du_1-du_2)(e_\alpha ) |^2 - \left( n{\widetilde{C}}_1 +C\right) \rho ^2\sum _{i=1}^2 |du_i(e_\alpha )|^2. \end{aligned} \end{aligned}$$

Namely,

$$\begin{aligned} \delta ^2(du_1, du_2) \ge \frac{n\lambda }{2} | du_1-du_2 |^2 - \left( n{\widetilde{C}}_1 +C\right) \rho ^2\sum _{i=1}^2 |du_i|^2. \end{aligned}$$
(3.1)

Therefore,

$$\begin{aligned} \begin{aligned}&\langle (\nabla h)\circ U, -T(dU(e_\alpha ), dU(e_\alpha ))\rangle \\&\quad \ge -s_\nu (\rho ) \left\{ \frac{n\sqrt{n\Lambda }}{\lambda } \rho \max | \nabla T| \sum _{i=1}^2 |du_i|^2 + \left( \frac{n{\widetilde{C}}_1}{\lambda }+\sqrt{C}\right) \rho \max |T|\sum _{i=1}^2 |du_i|^2 \right\} \\&\qquad -\frac{2n\varepsilon _1}{\lambda }\left[ \delta ^2(du_1, du_2)+ \left( n{\widetilde{C}}_1 +C\right) \rho ^2\sum _{i=1}^2 |du_i|^2\right] - \frac{n\Lambda }{2\varepsilon _1\lambda ^2}s_\nu (\rho )\rho \max |T|^2 |\sum _{i=1}^2du_i|^2 \\&\quad = - s_\nu (\rho )\rho \left\{ \frac{n\sqrt{n\Lambda }}{\lambda } \max | \nabla T| + \left( \frac{n{\widetilde{C}}_1}{\lambda }+\sqrt{C}\right) \max |T| + \frac{n\Lambda }{2\varepsilon _1\lambda ^2}\max |T|^2 \right\} \sum _{i=1}^2 |du_i|^2\\&\qquad -\frac{2n\varepsilon _1}{\lambda }\left[ \delta ^2(du_1, du_2)+ \left( n{\widetilde{C}}_1 +C\right) \rho ^2\sum _{i=1}^2 |du_i|^2\right] . \end{aligned} \end{aligned}$$

The above inequality and (2.3) imply that

$$\begin{aligned} \begin{aligned} \Delta _V \psi =&\Delta _V(h\circ U) = \sum _\alpha \nabla ^2 h(dU(e_\alpha ), dU(e_\alpha )) + \langle (\nabla h)\circ U, \tau _V(U) \rangle .\\ =&\sum _\alpha \nabla ^2 h(dU(e_\alpha ), dU(e_\alpha )) + \sum _\alpha \langle (\nabla h)\circ U, -T(dU(e_\alpha ), dU(e_\alpha ))\rangle \\ \ge&\left( s'_\nu (\rho )-\frac{2n\varepsilon _1}{\lambda }\right) \delta ^2(du_1, du_2) - s_\nu (\rho ) b_\kappa (\rho ) \sum _{i=1}^2 |du_i|^2 \\&- s_\nu (\rho )\rho \left\{ \frac{n\sqrt{n\Lambda }}{\lambda } \max | \nabla T| + \left( \frac{n{\widetilde{C}}_1}{\lambda }+\sqrt{C}\right) \max |T|\right. \\&\left. + \frac{n\Lambda }{2\varepsilon _1\lambda ^2}\max |T|^2 \right\} \sum _{i=1}^2 |du_i|^2\\&-\frac{2n\varepsilon _1}{\lambda }\left( n{\widetilde{C}}_1 +C\right) \rho ^2\sum _{i=1}^2 |du_i|^2. \end{aligned} \end{aligned}$$

Choosing \(\varepsilon _1= \frac{\lambda }{2n}\cos (\sqrt{\kappa }R)>0\), then we obtain

$$\begin{aligned} \begin{aligned} \Delta _V \psi \ge&- s_\nu (\rho ) b_\kappa (\rho ) \sum _{i=1}^2 |du_i|^2 - \cos (\sqrt{\kappa }R) \left( n{\widetilde{C}}_1 +C\right) \rho ^2\sum _{i=1}^2 |du_i|^2 \\&- s_\nu (\rho )\rho \left\{ \frac{n\sqrt{n\Lambda }}{\lambda } \max | \nabla T| + \left( \frac{n{\widetilde{C}}_1}{\lambda }+\sqrt{C}\right) \max |T| \right. \\&\left. + \frac{n^2\Lambda }{\lambda ^3\cos (\sqrt{\kappa }R)}\max |T|^2 \right\} \sum _{i=1}^2 |du_i|^2. \end{aligned} \end{aligned}$$
(3.2)

It follows from (2.1) that

$$\begin{aligned} \begin{aligned} \Delta _V \psi _i&= \Delta _V(\phi \circ U_i) = \sum _\alpha \nabla ^2 \phi (dU_i(e_\alpha ), dU_i(e_\alpha )) + \langle (\nabla \phi )\circ U_i, \tau _V(U_i) \rangle \\&\ge s_\kappa '(\rho _i)|du_i|^2 +s_\kappa (\rho _i) \langle (\nabla d)\circ U_i, 0\oplus (-T(du_i(e_\alpha ), du_i(e_\alpha ))) \rangle \\&\ge s_\kappa '(\rho _i)|du_i|^2 -s_\kappa (\rho _i) \left| T(du_i(e_\alpha ), du_i(e_\alpha )) \right| \\&\ge s_\kappa '(\rho _i)|du_i|^2 -s_\kappa (\rho _i) \max |T| |du_i|^2. \end{aligned} \end{aligned}$$
(3.3)

It is easy to check that

$$\begin{aligned} \omega ''=\omega '^2, \quad \quad \omega '\circ \psi _i=\frac{1}{q_\kappa ((1+\sigma )R)-q_\kappa (\rho _i)}. \end{aligned}$$

Therefore,

$$\begin{aligned} \begin{aligned}&{\mathcal {L}}_V (\Theta ) = \Delta _V\psi +\psi (\Delta _V\varPhi - |\nabla \varPhi |^2)\\&\quad = \Delta _V\psi + \frac{1}{2}\psi \sum _{i=1}^2 (\omega '\circ \psi _i)\Delta _V \psi _i + \sum _{i=1}^2 \left[ \frac{1}{2}\psi (\omega ''\circ \psi _i) - \frac{1}{4}\psi (\omega '\circ \psi _i)^2\right] |\nabla \psi _i|^2\\&\quad = \Delta _V\psi + \frac{1}{2}\psi \sum _{i=1}^2 (\omega '\circ \psi _i)\Delta _V \psi _i + \frac{\psi }{4}\sum _{i=1}^2(\omega '\circ \psi _i)^2|\nabla \psi _i|^2\\&\quad \ge -\, s_\nu (\rho ) b_\kappa (\rho ) \sum _{i=1}^2 |du_i|^2 - \cos (\sqrt{\kappa }R) \left( n{\widetilde{C}}_1 +C\right) \rho ^2\sum _{i=1}^2 |du_i|^2 \\&\qquad -\, s_\nu (\rho )\rho \left\{ \frac{n\sqrt{n\Lambda }}{\lambda } \max | \nabla T| + \left( \frac{n{\widetilde{C}}_1}{\lambda }+\sqrt{C}\right) \max |T|\right. \\&\left. \qquad +\, \frac{n^2\Lambda }{\lambda ^3\cos (\sqrt{\kappa }R)}\max |T|^2 \right\} \sum _{i=1}^2 |du_i|^2 \\&\qquad +\, \frac{1}{2}\psi \sum _{i=1}^2 \frac{1}{q_\kappa ((1+\sigma )R)-q_\kappa (\rho _i)}\left[ s_\kappa '(\rho _i)|du_i|^2 -s_\kappa (\rho _i) \max |T| |du_i|^2 \right] \\&\quad = \psi \sum _{i=1}^2 \left\{ \frac{s_{\kappa }'(\rho _i)}{2(q_\kappa ((1+\sigma )R)-q_\kappa (\rho _i))} - \frac{b_\kappa (\rho )s_\nu (\rho )}{q_\nu (\rho )} \right. \\&\left. \qquad -\, \frac{s_\nu (\rho )\rho }{q_\nu (\rho )}\left[ \frac{n\sqrt{n\Lambda }}{\lambda } \max | \nabla T| +\left( \frac{n{\widetilde{C}}_1}{\lambda }+\sqrt{C}\right) \max |T| \right. \right. \\&\qquad \left. +\, \frac{n^2\Lambda }{\lambda ^3\cos (\sqrt{\kappa }R)}\max |T|^2 \right] - \frac{s_\kappa (\rho _i)}{2(q_\kappa ((1+\sigma )R)-q_\kappa (\rho _i))}\max |T|\\&\left. \qquad - \frac{\rho ^2}{q_\nu (\rho )}\cos (\sqrt{\kappa }R) \left( n{\widetilde{C}}_1 +C\right) \right\} |du_i|^2 \\&\quad \ge \psi \sum _{i=1}^2 \left\{ \frac{s_{\kappa }'(\rho _i)}{2(q_\kappa ((1+\sigma )R)-q_\kappa (\rho _i))} - \frac{b_\kappa (\rho )s_\nu (\rho )}{q_\nu (\rho )}\right. \\&\left. \qquad - \left[ \frac{2n\sqrt{n\Lambda }}{\lambda } \max | \nabla T| +\left( \frac{2n{\widetilde{C}}_1}{\lambda }+2\sqrt{C}\right) \max |T| \right. \right. \\&\qquad \left. \left. + \frac{2n^2\Lambda }{\lambda ^3\cos (\sqrt{\kappa }R)}\max |T|^2 \right] - \frac{\kappa R^2}{1-\cos (\sqrt{\kappa }R)}\cos (\sqrt{\kappa }R) \left( n{\widetilde{C}}_1 +C\right) \right. \\&\qquad \left. - \frac{\sqrt{\kappa }\sin (\sqrt{\kappa }R)}{2\left[ \cos (\sqrt{\kappa }R)-\cos ((1+\sigma )\sqrt{\kappa }R)\right] }\max |T| \right\} |du_i|^2, \end{aligned} \end{aligned}$$
(3.4)

where we have used the fact that \(\frac{s_\nu (\rho )\rho }{q_\nu (\rho )}\) is nonincreasing in (0, 2R] and \(\frac{\rho ^2}{q_\nu (\rho )}\) is increasing in (0, 2R]. Direct computation gives us

$$\begin{aligned} \frac{b_\kappa (\rho )s_\nu (\rho )}{q_\nu (\rho )}= \frac{\kappa }{4}(1+\frac{1}{s'_{\frac{\kappa }{4}}(\rho )})(1+\frac{\rho }{s_\kappa (\rho )})=:\alpha (\rho ), \end{aligned}$$

and \(\alpha (t)\) is increasing in [0, 2R] and \(\beta _R(t)=\frac{s'_\kappa (t) }{2(q_\kappa ((1+\sigma )R)-q_\kappa (t))}\) is increasing in [0, R]. Hence, we obtain

$$\begin{aligned} \begin{aligned}&\frac{s_{\kappa }'(\rho _i)}{2(q_\kappa ((1+\sigma )R)-q_\kappa (\rho _i))} - \frac{b_\kappa (\rho )s_\nu (\rho )}{q_\nu (\rho )} \\&\quad \ge \frac{1}{2q_\kappa ((1+\sigma )R) }- \frac{\kappa }{4}\left( 1+\frac{1}{s'_{\frac{\kappa }{4}}(2R)}\right) \left( 1+\frac{2R}{s_\kappa (2R)}\right) \\&\quad = \frac{\kappa }{2[1-\cos ((1+\sigma )\sqrt{\kappa }R)]}- \frac{\kappa [1+\cos (\sqrt{\kappa }R)][2\sqrt{\kappa }R+\sin (2\sqrt{\kappa }R)]}{4\cos (\sqrt{\kappa }R)\sin (2\sqrt{\kappa }R)}. \end{aligned} \end{aligned}$$

It follows that

$$\begin{aligned} \begin{aligned} {\mathcal {L}}_V(\Theta )\ge&\psi \sum _{i=1}^2 \left\{ \frac{\kappa }{2[1-\cos ((1+\sigma )\sqrt{\kappa }R)]}- \frac{\kappa [1+\cos (\sqrt{\kappa }R)][2\sqrt{\kappa }R+\sin (2\sqrt{\kappa }R)]}{4\cos (\sqrt{\kappa }R)\sin (2\sqrt{\kappa }R)} \right. \\&\quad - \frac{2n\sqrt{n\Lambda }}{\lambda } \max | \nabla T|\\&\quad - \left( \frac{2n{\widetilde{C}}_1}{\lambda }+2\sqrt{C}+ \frac{\sqrt{\kappa }\sin (\sqrt{\kappa }R)}{2\left[ \cos (\sqrt{\kappa }R)-\cos ((1+\sigma )\sqrt{\kappa }R)\right] }\right) \max |T| \\&\quad \left. - \frac{2n^2\Lambda }{\lambda ^3\cos (\sqrt{\kappa }R)}\max |T|^2 - \frac{\left( n{\widetilde{C}}_1 +C\right) \kappa R^2\cos (\sqrt{\kappa }R)}{1-\cos (\sqrt{\kappa }R)} \right\} |du_i|^2.\\ \end{aligned} \end{aligned}$$

Clearly, by choosing appropriate \(\sigma \) and R, we obtain

$$\begin{aligned}&\frac{\kappa }{2[1-\cos ((1+\sigma )\sqrt{\kappa }R)]}- \frac{\kappa [1+\cos (\sqrt{\kappa }R)][2\sqrt{\kappa }R+\sin (2\sqrt{\kappa }R)]}{4\cos (\sqrt{\kappa }R)\sin (2\sqrt{\kappa }R)}\\&\quad - \frac{\left( n{\widetilde{C}}_1 +C\right) \kappa R^2\cos (\sqrt{\kappa }R)}{1-\cos (\sqrt{\kappa }R)} > 0. \end{aligned}$$

Hence, if

$$\begin{aligned} \begin{aligned}&\frac{2n\sqrt{n\Lambda }}{\lambda } \max | \nabla T| +\left( \frac{2n{\widetilde{C}}_1}{\lambda }+2\sqrt{C}+ \frac{\sqrt{\kappa }\sin (\sqrt{\kappa }R)}{2\left[ \cos (\sqrt{\kappa }R)-\cos ((1+\sigma )\sqrt{\kappa }R)\right] }\right) \max |T|\\&\qquad + \frac{2n^2\Lambda }{\lambda ^3\cos (\sqrt{\kappa }R)}\max |T|^2 \\&\quad \le \frac{\kappa }{2[1-\cos ((1+\sigma )\sqrt{\kappa }R)]}- \frac{\kappa [1+\cos (\sqrt{\kappa }R)][2\sqrt{\kappa }R+\sin (2\sqrt{\kappa }R)]}{4\cos (\sqrt{\kappa }R)\sin (2\sqrt{\kappa }R)}\\&\qquad - \frac{\left( n{\widetilde{C}}_1 +C\right) \kappa R^2\cos (\sqrt{\kappa }R)}{1-\cos (\sqrt{\kappa }R)}, \end{aligned} \end{aligned}$$
(3.5)

then we have

$$\begin{aligned} {\mathcal {L}}_V(\Theta ) \ge 0. \end{aligned}$$

(For \(\sqrt{\kappa }R \rightarrow 0\), we use the Taylor expansions of \(\sin \) and \(\cos \) to obtain positive values on the right-hand side of (3.5).) It is easy to see that there exists a constant \(C_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, so that if

$$\begin{aligned} \max |\nabla T| + \max |T| \le C_0, \end{aligned}$$
(3.6)

then (3.5) holds true; consequently, \({\mathcal {L}}_V(\Theta ) \ge 0\). Applying the ordinary maximum principle, we have

$$\begin{aligned} \max _M \Theta \le \max _{\partial M}\Theta . \end{aligned}$$

\(\square \)

Proof of Theorem 2

We consider a parabolic operator of the form

$$\begin{aligned} \widetilde{{\mathcal {L}}}_V := {\mathcal {L}}_V - e^{-\varPhi }\partial _t. \end{aligned}$$

By using

$$\begin{aligned} \left( \Delta _V -\partial _t\right) \psi= & {} \left( \Delta _V-\partial _t\right) (h\circ U) = \sum _\alpha \nabla ^2 h(dU(e_\alpha ), dU(e_\alpha ))\\&\quad + \langle (\nabla h)\circ U, \tau _V(U) -\partial _t U \rangle \end{aligned}$$

and

$$\begin{aligned} \left( \Delta _V-\partial _t\right) \psi _i= & {} \left( \Delta _V-\partial _t\right) (\phi \circ U_i) = \sum _\alpha \nabla ^2 \phi (dU_i(e_\alpha ), dU_i(e_\alpha )) \\&\quad + \langle (\nabla \phi )\circ U_i, \tau _V(U_i) - \partial _t U_i \rangle , \end{aligned}$$

as in the proof of Theorem 1, we can conclude that \(\widetilde{{\mathcal {L}}}_V(\Theta )\ge 0\) on \(M_T\). From the parabolic maximum principle, we have

$$\begin{aligned} \max _{M_T} \Theta \le \max _{\partial _p M_T}\Theta . \end{aligned}$$

4 Existence results

Using the maximum principle obtained in the last section, we shall prove the existence of solutions of the Dirichlet problem for \(VT-\)harmonic maps.

Proof of Theorem 3

Let us choose normal coordinates \(\{ y^i \}_{i=1,2,\ldots ,n}\) centered at p, then any \(VT-\)harmonic map \(u: M\rightarrow B_R(p)\subset N\) can be written as

$$\begin{aligned} u=(u^1, \ldots , u^n) \in (H^{2, q}(M))^n \end{aligned}$$

which satisfies the elliptic system

$$\begin{aligned} \Delta _M u^i + \Gamma ^{i}_{jk}(u)\frac{\partial u^j}{\partial x^\alpha }\frac{\partial u^k}{\partial x^\beta }g^{\alpha \beta } + V^\alpha \frac{\partial u^i}{\partial x^\alpha } + T^{i}_{jk}(u)\frac{\partial u^j}{\partial x^\alpha } \frac{\partial u^k}{\partial x^\beta }g^{\alpha \beta } = 0, \ \ \ i=1, 2, \ldots , n. \end{aligned}$$

For simplicity of notation, we write it in a concise form

$$\begin{aligned} \Delta u+ \Gamma (du, du) +du(V) + \mathrm{Tr}_g T(du, du) = 0. \end{aligned}$$

Now we consider the initial boundary value problem for the heat flow of \(VT-\)harmonic maps

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle \partial _t u = \Delta u + \Gamma (du, du) + du(V) + \mathrm{Tr}_g T(du, du), \\ \displaystyle u-u_0 \in H^{2, q}_0(M, N), \quad u(0)=u_0, \\ \displaystyle u(M\times [0, +\infty )) \subset B_R(p). \end{array} \right. \end{aligned}$$
(4.1)

As in the proof of Theorem 3 in [5], by a continuity method that rests on the maximum principle Theorem 2, we can conclude the global existence of a solution u(xt) of the above flow (4.1). This solution satisfies

$$\begin{aligned} \Vert u(\cdot , t) \Vert _{1+\alpha } \le C(q, M, V, T, N, u_0, R), \quad \forall t\in (0, +\infty ) \end{aligned}$$

for some \(\alpha >0\). Consequently, by the parabolic regularity theory, we have the uniform estimate

$$\begin{aligned} \Vert u\Vert _{C^{1+\alpha , 2+\alpha }(M)} \le C. \end{aligned}$$
(4.2)

For \(u_1(x, t) = u(x, t), u_2(x, t) = u(x, t+\sigma _1), \sigma _1>0, \forall (x, t)\in M\times (0, +\infty )\), as in the proof of Theorem 2, the function \(\Theta \) satisfies

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle (\Delta -\partial _t)(\frac{\Theta }{\sigma _{1}^2}) + \langle V-2\nabla \varPhi , \nabla (\frac{\Theta }{\sigma _{1}^2}) \rangle \ge 0, \\ \displaystyle \left. \Theta \right| _{\partial M} = 0. \end{array} \right. \end{aligned}$$

By the ordinary maximum principle for functions, it follows that (see pp.178–179 in [17])

$$\begin{aligned} \left( \frac{\Theta }{\sigma _{1}^2}\right) \le C(t-t_0)^{-k}, \quad \forall t\ge t_0 \end{aligned}$$

for any positive integer k and some \(t_0>0\). Letting \(\sigma _1 \rightarrow 0\), then we obtain \(|u_t| \rightarrow 0\) as \(t\rightarrow +\infty \), from which together with (4.2), we have u subconverges to a \(VT-\)harmonic map \(u_\infty \) satisfying (1.7) and \(u_\infty (M)\subset B_R(p)\). \(\square \)

With the Schauder and higher regularity estimates, we can improve Theorem 3 to the following

Theorem 5

Let \(M, N, V, T, B_R(p)\) be as in Theorem 1. Suppose \(u_0 \in C^0(M, N)\) with \(u_0(M) \subset B_R(p)\). For appropriate \(\sigma \) and R, there exists a constant \(C_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla T| + \max |T| \le C_0, \end{aligned}$$
(4.3)

then the Dirichlet problem

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle \tau (u) + du(V) + \mathrm{Tr}_g T(du, du) = 0, \\ \displaystyle \left. u\right| _{\partial M}= \left. u_0\right| _{\partial M} \end{array} \right. \end{aligned}$$

admits a unique solution \(u \in C^\infty (M, N)\cap C^0({\overline{M}}, N)\) such that \(u(M)\subset B_R(p)\).

5 Applications

5.1 Weyl harmonic maps (c.f. [14])

Let \((M, [g], {}^W\nabla )\) be a Weyl manifold. According to the definition, there exists a 1-form \(\Theta \) such that \({}^W g= \Theta \otimes g\) for any \(g \in [g]\). Equivalently, \({}^W\nabla \) is defined by

$$\begin{aligned} {}^W\nabla _{X}Y = \nabla _X Y - \frac{1}{2}\Theta (X)Y - \frac{1}{2}\Theta (Y)X + \frac{1}{2}g(X, Y)\Theta ^{\sharp }, \ \ \ \forall X, Y \in \Gamma (TM), \end{aligned}$$

where \(\nabla \) is the Levi-Civita connection and \(\Theta ^\sharp \) the vector field dual to \(\Theta \) w.r.t. g. Let \(\Gamma ^\gamma _{\alpha \beta }, {}^{W}\Gamma ^{\gamma }_{\alpha \beta }\) be the Christoffel symbols corresponding to \(\nabla \) and \({}^W\nabla \), respectively.

Let \((N, [{\tilde{g}}], {}^W{\tilde{\nabla }})\) be also a Weyl manifold, and correspondingly, we denote by \({\widetilde{\Theta }}\) the 1-form, and \(\Gamma ^{k}_{ij}, {}^{W}{\tilde{\Gamma }}^{k}_{ij}\) are the Christoffel symbols for the Levi-Civita connection \({\tilde{\nabla }}\) and Weyl connection \({}^W{\tilde{\nabla }}\), respectively. Let \(u: (M, [g], {}^W\nabla ) \rightarrow (N, [{\tilde{g}}], {}^W{\tilde{\nabla }})\) be the usual smooth map.

Let

$$\begin{aligned}\begin{aligned} V:=&(\Gamma ^\gamma _{\alpha \beta } - {}^W\Gamma ^\gamma _{\alpha \beta } )g^{\alpha \beta }\frac{\partial }{\partial x^\gamma },\\ T({\tilde{X}}, {\tilde{Y}}) :=&- \frac{1}{2}{\widetilde{\Theta }}({\tilde{X}}){\tilde{Y}} - \frac{1}{2}{\widetilde{\Theta }}({\tilde{Y}}){\tilde{X}} + \frac{1}{2}g({\tilde{X}}, {\tilde{Y}}){\widetilde{\Theta }}^{\sharp }, \ \ \ \forall {\tilde{X}}, {\tilde{Y}} \in \Gamma (TN). \end{aligned} \end{aligned}$$

Then, we have

$$\begin{aligned} \begin{aligned} \tau (g, {}^W\nabla , {}^W{\tilde{\nabla }}) =&g^{\alpha \beta }( u^{k}_{\alpha \beta }+ {}^{W}{\tilde{\Gamma }}^{k}_{ij}u^{i}_\alpha u^{j}_\beta - {}^W\Gamma ^\sigma _{\alpha \beta }u^{k}_\sigma )\partial _{y^k} \\ =&[g^{\alpha \beta }( u^{k}_{\alpha \beta }+ \Gamma ^{k}_{ij}u^{i}_\alpha u^{j}_\beta +T^{k}_{ij}u^{i}_\alpha u^{j}_\beta - \Gamma ^{\sigma }_{\alpha \beta }u^{k}_\sigma )+ V^\sigma u^{k}_\sigma ]\partial _{y^k} \\ =&\tau (g, \nabla , {\tilde{\nabla }}) +du(V) + \mathrm{Tr}_g (du, du). \end{aligned} \end{aligned}$$

Hence,

$$\begin{aligned} \tau (g, {}^W\nabla , {}^W{\tilde{\nabla }}) =0 \quad \mathrm{iff} \quad \tau (u) + du(V) + \mathrm{Tr}_g T(du, du) = 0. \end{aligned}$$

Corollary 1

Let \((M, [g], {}^W\nabla )\) be a compact Wey manifold with nonempty boundary \(\partial M\) and \((N, [{\tilde{g}}], {}^W{\tilde{\nabla }})\) a complete Weyl manifold with sectional curvature bounded from above by \(\kappa \ge 0\). Let \(u_0: M \rightarrow N\) be a continuous map with \(u_0(M) \subset B_R(p)\), a geodesic ball with radius \(R< \frac{\pi }{2(1+\sigma )\sqrt{\kappa }}\). For appropriate \(\sigma \) and R, there exists a constant \(C_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla {\widetilde{\Theta }}| + \max |{\widetilde{\Theta }}| \le C_0, \end{aligned}$$

then there exists a unique Weyl harmonic map \(u: M\rightarrow B_R(p) \subset N\) with \(u=u_0\) on \(\partial M\).

5.2 Affine harmonic maps (c.f. [9, 10])

Let \((M, g, {\tilde{\nabla }}), (N, h, {\tilde{\nabla }}')\) both be affine manifolds, where \({\tilde{\nabla }}\) is a global flat and torsion-free connection on M and \({\tilde{\nabla }}'\) is a torsion-free connection on N. Then, we have

$$\begin{aligned} \tau (g, {\tilde{\nabla }}, {\tilde{\nabla }}') = g^{\alpha \beta }(u^{k}_{\alpha \beta }+{\tilde{\Gamma }}'^{k}_{ij}u^{i}_{\alpha }u^{j}_\beta )\partial _{y^k}, \end{aligned}$$

where \( {\tilde{\Gamma }}'^{k}_{ij}\) are the Christoffel symbols of \({\tilde{\nabla }}'\).

Regarding (Mg) and (Nh) as Riemannian manifolds, let \(\Gamma ^{\gamma }_{\alpha \beta }\) and \(\Gamma ^{i}_{jk}\) be the Christoffel symbols of the Levi-Civita connections \(\nabla \) and \(\nabla '\) of (Mg) and (Nh), respectively. We then have the usual tension field

$$\begin{aligned} \tau (g, \nabla , \nabla ' ) = g^{\alpha \beta } (u^{k}_{\alpha \beta } - \Gamma ^{\gamma }_{\alpha \beta }u^{k}_\gamma + \Gamma ^{k}_{ij}u^{i}_\alpha u^{j}_\beta )\partial _{y^k}. \end{aligned}$$

Let

$$\begin{aligned} V:= & {} g^{\alpha \beta }\Gamma ^{\gamma }_{\alpha \beta }\partial _{x^\gamma }, \\ T^{k}_{ij}:= & {} {\tilde{\Gamma }}'^{k}_{ij} - \Gamma ^{k}_{ij}. \end{aligned}$$

Then, we have

$$\begin{aligned} \begin{aligned} \tau (g, {\tilde{\nabla }}, {\tilde{\nabla }}') =\tau (g, \nabla , \nabla ' ) +du(V) + \mathrm{Tr}_g (du, du). \end{aligned} \end{aligned}$$

Therefore,

$$\begin{aligned} \tau (g, {\tilde{\nabla }}, {\tilde{\nabla }}') =0 \quad \mathrm{iff} \quad \tau (u) + du(V) + \mathrm{Tr}_g T(du, du) = 0. \end{aligned}$$

Corollary 2

Let \((M, g, {\tilde{\nabla }})\) be a compact affine manifold with nonempty boundary \(\partial M\) and \((N, h, {\tilde{\nabla }}')\) a complete affine manifold with sectional curvature bounded from above by \(\kappa \ge 0\), where \({\tilde{\nabla }}\) is a global flat and torsion-free connection on M and \({\tilde{\nabla }}'\) is a torsion-free connection on N. Let \(u_0: M \rightarrow N\) be a continuous map with \(u_0(M) \subset B_R(p)\), a geodesic ball with radius \(R< \frac{\pi }{2(1+\sigma )\sqrt{\kappa }}\). Denote \(T^{k}_{ij}:= {\tilde{\Gamma }}'^{k}_{ij} - \Gamma ^{k}_{ij}\), where \( {\tilde{\Gamma }}'^{k}_{ij}\) and \(\Gamma ^{i}_{jk}\) stand for the Christoffel symbols of \({\tilde{\nabla }}'\) and \(\nabla '\), respectively. For appropriate \(\sigma \) and R, there exists a constant \(C_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla T| + \max |T|\le C_0, \end{aligned}$$

then there exists a unique affine harmonic map \(u: M\rightarrow B_R(p) \subset N\) with \(u=u_0\) on \(\partial M\).

5.3 Hermitian harmonic maps (c.f. [11, 15])

Let \((M^m, g, {\tilde{\nabla }}), (N^n, h, {\tilde{\nabla }}')\) are both Hermitian manifolds, where \({\tilde{\nabla }}\) and \({\tilde{\nabla }}'\) are holomorphic torsion-free connections on M and N, respectively. Direct calculation gives us

$$\begin{aligned} \tau (g, {\tilde{\nabla }}, {\tilde{\nabla }}') = g^{\alpha {\bar{\beta }}}\left( \frac{\partial ^2 u^i}{\partial z^\alpha \partial {\bar{z}}^\beta } + \Gamma '^{i}_{jk}\frac{\partial u^{j}}{\partial z^\alpha }\frac{\partial u^k}{\partial {\bar{z}}^\beta } \right) \frac{\partial }{\partial w^i} + \overline{g^{\alpha {\bar{\beta }}}\left( \frac{\partial ^2 u^i}{\partial z^\alpha \partial {\bar{z}}^\beta } + \Gamma '^{i}_{jk}\frac{\partial u^{j}}{\partial z^\alpha }\frac{\partial u^k}{\partial {\bar{z}}^\beta } \right) }\frac{\partial }{\partial {\bar{w}}^i}, \end{aligned}$$

where \( \Gamma '^{i}_{jk}\) are the Christoffel symbols of \({\tilde{\nabla }}'\).

Let J be the almost complex structure, and \(\{e_A\}=\{e_1, \ldots , e_m, Je_1, \ldots , Je_m\}\) a local basis of M. Let \(\nabla , \nabla '\) be the Levi-Civita connections on M and N, respectively, and \(\Gamma ^{i}_{jk}\)the Christoffel symbols of \(\nabla '\). Set

$$\begin{aligned} V:= {\tilde{\nabla }}_{e_A}e_A - \nabla _{e_A}e_A \quad \mathrm{and} \quad T^{i}_{jk} := \Gamma '^{i}_{jk} - \Gamma ^{i}_{jk}, \end{aligned}$$

then we have

$$\begin{aligned} \tau (g, {\tilde{\nabla }}, {\tilde{\nabla }}') = \tau (u) + du(V) + \mathrm{Tr}_g T(du, du). \end{aligned}$$

Namely,

$$\begin{aligned} \tau (g, {\tilde{\nabla }}, {\tilde{\nabla }}') =0 \quad \mathrm{iff} \quad \tau (u) + du(V) + \mathrm{Tr}_g T(du, du) = 0. \end{aligned}$$

Corollary 3

Let \((M^m, g, {\tilde{\nabla }})\) be a compact Hermitian manifold with nonempty boundary \(\partial M\) and \((N^n, h, {\tilde{\nabla }}')\) a complete Hermitian manifold with sectional curvature bounded from above by \(\kappa \ge 0\), where \({\tilde{\nabla }}\) and \({\tilde{\nabla }}'\) are holomorphic torsion-free connections on M and N, respectively. Let \(u_0: M \rightarrow N\) be a continuous map with \(u_0(M) \subset B_R(p)\), a geodesic ball with radius \(R< \frac{\pi }{2(1+\sigma )\sqrt{\kappa }}\). Denote \(T^{k}_{ij}:= \Gamma '^{k}_{ij} - \Gamma ^{k}_{ij}\), where \( \Gamma '^{k}_{ij}\) and \(\Gamma ^{i}_{jk}\) stand for the Christoffel symbols of \({\tilde{\nabla }}'\) and \(\nabla '\), respectively. For appropriate \(\sigma \) and R, there exists a constant \(C_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla T| + \max |T| \le C_0, \end{aligned}$$

then there exists a unique Hermitian harmonic map \(u: M\rightarrow B_R(p) \subset N\) with \(u=u_0\) on \(\partial M\).

5.4 Magnetic harmonic maps

We now consider a case that, in contrast to the previous ones, does not arise from a structure different from the Riemannian, but from on additional structure on a Riemannian manifold. Let \( (\Sigma ^m, g)\) be an m-dimensional compact oriented Riemannian manifold with nonempty boundary, \((N, {\widetilde{g}})\) a Riemannian manifold of dimension n. Let \(u: (\Sigma ^m, g)\rightarrow (N, {\widetilde{g}})\) be a map and \(Z\in \Gamma (\mathrm{Hom}(\Lambda ^mTN, TN))\cong \Gamma (\Lambda ^mT^*N\otimes TN)\).

Consider the following system:

$$\begin{aligned} \tau (u) + Z(du(e_1) \wedge \cdots \wedge du(e_m)) = 0, \end{aligned}$$
(5.1)

where \(\{ e_1, \ldots , e_m \}\) is a positively oriented local orthonormal frame of \(\Sigma ^m\). In string theory, it can be interpreted as the motion equation of an \((m-1)\)-brane under an extrinsic magnetic force (c.f. [13]). In [13], the author obtained the global existence of the heat flow in one- dimensional case.

Using a similar method as above, in the two-dimensional case, we can obtain the following

Theorem 6

Let \(u_1, u_2 \in C^0(\Sigma ^2, N)\) be two magnetic harmonic maps into a geodesic ball \(B_{R}(p)\). For appropriate \(\sigma \) and R, there exists a constant \(C_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla Z| + \max |Z| \le C_0, \end{aligned}$$

then the function \(\Theta : \Sigma ^2 \rightarrow {\mathbb {R}}\) defined by

$$\begin{aligned} \Theta := \frac{q_{\frac{\kappa }{4}}(\rho )}{(q_\kappa ((1+\sigma )R)-q_\kappa (\rho _1))^{\frac{1}{2}}\cdot (q_\kappa ((1+\sigma )R)-q_\kappa (\rho _2))^{\frac{1}{2}}} \end{aligned}$$
(5.2)

satisfies the maximum principle, namely

$$\begin{aligned} \max _{\Sigma ^2} \Theta \le \max _{\partial \Sigma ^2}\Theta . \end{aligned}$$

Here \(\rho := d(u_1, u_2), \rho _i:= d(p, u_i), i=1,2.\)

In particular, if \(u_1=u_2\) on the boundary \(\partial \Sigma ^2\), then \(u_1\equiv u_2\) on \(\Sigma ^2\).

For the heat flow of magnetic harmonic maps, an analogous result holds. For \(T>0\), we set

$$\begin{aligned} \Sigma ^2_T:= \Sigma ^2\times [0, T] \end{aligned}$$

and denote the parabolic boundary of \(\Sigma ^2_T\) by

$$\begin{aligned} \partial _p \Sigma ^2_T:=(\Sigma ^2\times \{0\})\cup (\partial \Sigma ^2\times [0, T]). \end{aligned}$$

For the heat flow of magnetic harmonic maps

$$\begin{aligned} \partial _t u= \tau (u)+Z(du(e_1)\wedge du(e_2)), \end{aligned}$$
(5.3)

we have

Theorem 7

Let \(u_1, u_2 \in C^0(\Sigma ^2, N)\) be two solutions of heat flow Eq. (5.3) for magnetic harmonic maps into a geodesic ball \(B_{R}(p)\). For appropriate \(\sigma \) and R, there exists a constant \(C_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla Z| + \max |Z| \le C_0, \end{aligned}$$

then the function \(\Theta : \Sigma ^2_T \rightarrow {\mathbb {R}}\) defined by (5.2) with \(\Sigma ^2\) replaced by \(\Sigma ^2_T\) satisfies the maximum principle:

$$\begin{aligned} \max _{\Sigma ^2_T} \Theta \le \max _{\partial _p \Sigma ^2_T}\Theta . \end{aligned}$$

In particular, if \(u_1=u_2\) on the boundary \(\partial _p \Sigma ^2_T\), then \(u_1\equiv u_2\) on \(\Sigma ^2_T\).

As an application of the above maximum principle, we obtain the existence of magnetic harmonic maps into a geodesic ball.

Theorem 8

Let \(\Sigma ^2, N, Z, B_R(p)\) be as in Theorem 7. Suppose \(u_0 \in H^{2, q}(\Sigma ^2, N)(q>2)\) with \(u_0(\Sigma ^2) \subset B_R(p)\). For appropriate \(\sigma \) and R, there exists a constant \(C_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla Z| + \max |Z| \le C_0, \end{aligned}$$

then the initial boundary value problem

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle \partial _tu= \tau (u) +Z(du(e_1)\wedge du(e_2)), \\ \displaystyle u-u_0 \in H^{2, q}_0(\Sigma ^2, N),\quad u(0)=u_0, \quad u(\Sigma ^2\times [0,\infty )) \subset B_R(p), \end{array} \right. \end{aligned}$$
(5.4)

admits a unique global solution u which subconverges to a unique solution \(u \in H^{2, q}(\Sigma ^2, N)\) of the Dirichlet problem

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle \tau (u) + Z(du(e_1)\wedge du(e_2)) = 0, \\ \displaystyle u-u_0 \in H^{2, q}_0(\Sigma ^2, N), \end{array} \right. \end{aligned}$$
(5.5)

such that \(u(\Sigma ^2)\subset B_R(p)\).

Remark 2

In local coordinates \(\{ x^\alpha \}\) on M and \(\{ y^j \}\) on N, respectively, the term \(\mathrm{Tr}_g T(du, du)\) can be written as \(T^i_{jk}(u) \frac{\partial u^j}{\partial x^\alpha }\frac{\partial u^k}{\partial x^\beta }g^{\alpha \beta }\), where \(T:=T^i_{jk}\frac{\partial }{\partial y^i}\otimes dy^j \otimes dy^k\). Correspondingly, the term \(Z(du(e_1)\wedge du(e_2))\) in (5.3) could be written as \((Z^i_{jk}-Z^i_{kj})\frac{\partial u^j}{\partial x^\alpha }\frac{\partial u^k}{\partial x^\beta }g^{\alpha \beta }\). Using \(Z^i_{jk}-Z^i_{kj}\) in place of \(T^i_{jk}\) in the proof of the results for VT-harmonic maps, we can conclude the above theorems for magnetic harmonic maps.

6 VT-harmonic maps from complete manifolds into geodesic balls

In this section, we shall establish the existence of VT-harmonic maps from complete noncompact manifolds into geodesic balls in complete Riemannian manifolds with sectional curvature bounded above by a positive constant.

Before proving the existence theorem, we first give the following Bochner formula:

Lemma 3

Let \((M^m, g)\) and \((N^n, {\tilde{g}})\) be Riemannian manifolds. Let \(\mathrm{Ric}_V:= \mathrm{Ric} - \frac{1}{2}L_V g\), where \(\mathrm{Ric}\) is the Ricci curvature of M and \(L_V\) is the Lie derivative. Suppose u is a VT-harmonic map from M to N, then

$$\begin{aligned} \begin{aligned} \frac{1}{2}\Delta _V e(u) =&|\nabla du|^2 + \sum _{\alpha =1}^m\langle du(\mathrm{Ric}_V (e_\alpha )), du(e_\alpha ) \rangle \\&- \sum _{\alpha , \beta =1}^m R^N(du(e_\alpha ), du(e_\beta ), du(e_\alpha ), du(e_\beta ))\\&- \sum _{\alpha , \beta =1}^m \langle (\nabla _{e_\alpha }T)(du(e_\beta ), du(e_\beta )), du(e_\alpha )\rangle \\&- \sum _{\alpha , \beta =1}^m \langle 2T((\nabla _{e_\alpha }du)(e_\beta ), du(e_\beta )), du(e_\alpha ) \rangle , \end{aligned} \end{aligned}$$
(6.1)

where \(\{e_\alpha \}\) is a local orthonormal frame of M.

Proof

By the proof of Proposition 1.3.5 in [18] (c.f. also in [1]), we have

$$\begin{aligned} \begin{aligned} \frac{1}{2}\Delta e(u) =&|\nabla du|^2 + \sum _{\alpha =1}^m \langle \nabla _{e_\alpha }\tau (u), du(e_\alpha ) \rangle + \sum _{\alpha =1}^m\langle du(\mathrm{Ric} (e_\alpha )), du(e_\alpha ) \rangle \\&- \sum _{\alpha , \beta =1}^m R^N(du(e_\alpha ), du(e_\beta ), du(e_\alpha ), du(e_\beta )). \end{aligned} \end{aligned}$$

Let \(\{e_\alpha \}\) be a local orthonormal normal frame of M at the considered point. Since

$$\begin{aligned} \begin{aligned}&\sum _{\alpha =1}^m\langle \nabla _{e_\alpha }du(V), du(e_\alpha ) \rangle = \sum _{\alpha =1}^m\langle (\nabla _{e_\alpha }du)(V)+ du(\nabla _{e_\alpha }V), du(e_\alpha ) \rangle \\&\quad = \sum _{\alpha =1}^m\langle (\nabla _{V}du)(e_\alpha )+ du(\nabla _{e_\alpha }V), du(e_\alpha ) \rangle = \sum _{\alpha =1}^m \langle \nabla _V du(e_\alpha ) + du(\nabla _{e_\alpha }V), du(e_\alpha ) \rangle \\&\quad = \frac{1}{2}V|du|^2 + \sum _{\alpha , \beta =1}^m \langle \nabla _{e_\alpha }V, e_\beta \rangle \langle du(e_\beta ), du(e_\alpha ) \rangle \\&\quad = \frac{1}{2}V|du|^2 + \frac{1}{2}\sum _{\alpha , \beta =1}^m (L_V g)(e_\alpha , e_\beta )\langle du(e_\alpha , du(e_\beta )) \rangle \end{aligned}, \end{aligned}$$

and

$$\begin{aligned} \begin{aligned}&\nabla _{e_\alpha }(\mathrm{Tr}_g T(du,du)) = \nabla _{e_\alpha } (T(du(e_\beta ), du(e_\beta ))) \\&\quad = (\nabla _{e_\alpha }T)(du(e_\beta ), du(e_\beta ))+ 2T((\nabla _{e_\alpha }du)(e_\beta ), du(e_\beta )). \end{aligned} \end{aligned}$$

Therefore, we get

$$\begin{aligned} \begin{aligned} \frac{1}{2}\Delta e(u) =&|\nabla du|^2 - \frac{1}{2}V|du|^2 - \frac{1}{2}\sum _{\alpha , \beta =1}^m (L_V g)(e_\alpha , e_\beta )\langle du(e_\alpha , du(e_\beta )) \rangle \\&- \sum _{\alpha ,\beta =1}^m \langle (\nabla _{e_\alpha }T)(du(e_\beta ), du(e_\beta ))+ 2T((\nabla _{e_\alpha }du)(e_\beta ), du(e_\beta )), du(e_\alpha ) \rangle \\&+ \sum _{\alpha =1}^m\langle du(\mathrm{Ric} (e_\alpha )), du(e_\alpha ) \rangle - \sum _{\alpha , \beta =1}^m R^N(du(e_\alpha ), du(e_\beta ), du(e_\alpha ), du(e_\beta )), \end{aligned} \end{aligned}$$

which implies that (6.1) holds. \(\square \)

Using the above Bochner formula and the estimate of \(\Delta _V r\) in [6] (here r denotes the distance function on M), we establish the gradient estimate for VT-harmonic maps.

Theorem 9

Let \((M^m, g)\) be a complete noncompact Riemannian manifold with

$$\begin{aligned} \mathrm{Ric}_V:= \mathrm{Ric} - \frac{1}{2}L_Vg \ge -A, \end{aligned}$$

where \(A \ge 0\) is a constant, \(\mathrm{Ric}\) is the Ricci curvature of M and \(L_V\) is the Lie derivative. Let \((N^n, {\widetilde{g}})\) be a complete Riemannian manifold with sectional curvature bounded from above by a positive constant \(\kappa \). Let \(u: M\rightarrow N\) be a VT-harmonic map such that \(u(M) \subset B_{{\widetilde{R}}}(p)\), where \(B_{{\widetilde{R}}}(p)\) is a regular ball in N, i.e., disjoint from the cut locus of p and \({\widetilde{R}} < \frac{\pi }{2\sqrt{\kappa }}\). Suppose \(\Vert V\Vert _{L^\infty }< +\infty , \Vert T \Vert _{L^\infty }<+\infty , \Vert \nabla T \Vert _{L^\infty }<+\infty \) and

$$\begin{aligned} \left( 1+(m+1)^2 - \frac{1}{(m+1)^2} \right) \Vert T\Vert ^2_{L^\infty } + \frac{\sqrt{\kappa }}{\cos (\sqrt{\kappa }\widetilde{R)}}\Vert T\Vert _{L^\infty } < \frac{\kappa }{\min \{m, n\}}. \end{aligned}$$
(6.2)

Then, we have

$$\begin{aligned} |\nabla u|\le C_6(\sqrt{A} + 1 ), \end{aligned}$$

where \(C_6>0\) is a constant depending only on \(m, n, \kappa , {\tilde{R}}, V, T\).

Proof

Let \(r, \rho \) be the respective distance functions on M and N from some fixed points \({\tilde{p}}\in M, p \in N\). Let \(B_a({\tilde{p}})\) be a geodesic ball of radius a around \({\tilde{p}}\). Define \(\varphi :=\cos (\sqrt{\kappa }\rho )\). Then, the Hessian comparison theorem implies

$$\begin{aligned} \mathrm{Hess}^N(\varphi ) \le -\kappa (\cos (\sqrt{\kappa }\rho )){\widetilde{g}}. \end{aligned}$$
(6.3)

Define \(f: B_a({\tilde{p}}) \rightarrow {\mathbb {R}}\) by

$$\begin{aligned} f:= (a^2-r^2)\frac{|\nabla u|}{\varphi \circ u}. \end{aligned}$$

Denote \(\psi := \frac{|\nabla u|}{\varphi \circ u}\). Clearly, f achieves its maximum at some interior point of \(B_a({\tilde{p}})\), say q. WLOG, we assume that \(\nabla u(q) \ne 0\). Then, from

$$\begin{aligned} \begin{aligned} \nabla f(q) =&0, \\ \Delta _V f(q) \le&0, \end{aligned} \end{aligned}$$

we obtain at q:

$$\begin{aligned}&\frac{\nabla r^2}{a^2-r^2}=\frac{\nabla \psi }{\psi }, \end{aligned}$$
(6.4)
$$\begin{aligned}&\frac{\Delta _V\psi }{\psi } - \frac{\Delta _V r^2}{a^2-r^2} - \frac{2\langle \nabla r^2, \nabla \psi \rangle }{(a^2-r^2)\psi }\le 0. \end{aligned}$$
(6.5)

It follows from the above two inequalities that

$$\begin{aligned} \frac{\Delta _V\psi }{\psi } - \frac{\Delta _V r^2}{a^2-r^2} - \frac{2|\nabla r^2|^2}{(a^2-r^2)^2}\le 0. \end{aligned}$$
(6.6)

By formula (2.4) in [6] (see also [16]), we have

$$\begin{aligned} \Delta _Vr^2 = 2r\Delta _V r + 2|\nabla r|^2 \le 2r(A(r-r_0)+{\widetilde{C}}_0)+2, \end{aligned}$$
(6.7)

where \(r_0>0\) is a sufficiently small constant and \({\widetilde{C}}_0 := \max _{\partial B_{r_0}({\tilde{p}})}\Delta _V r\).

Let \(\{e_\alpha \}\) be a local orthonormal frame field of M and s the rank of u at the point. We shall compute in normal coordinates at the considered point of N. By Newton’s inequality, we get

$$\begin{aligned} \begin{aligned}&\sum _{\alpha , \beta }R^N(du(e_\alpha ), du(e_\beta ), du(e_\alpha ), du(e_\beta )) = \sum _{\alpha \ne \beta }R^N(du(e_\alpha ), du(e_\beta ), du(e_\alpha ), du(e_\beta ))\\&\quad \le 2\kappa \sum _{1\le \alpha < \beta \le s}\left( \sum _i(u^i_{\alpha })^2\right) \left( \sum _j(u^j_{\beta })^2\right) \le 2 \kappa \cdot \frac{s(s-1)}{2} \cdot \frac{1}{s^2} \left( \sum _{\alpha , i} (u^i_{\alpha })^2 \right) ^2 \\&\quad = \frac{s-1}{s}\kappa |du|^4 \le \frac{s_0-1}{s_0}\kappa |du|^4, \end{aligned} \end{aligned}$$

where we have used the fact that \(s_0:=\min \{m, n\}\ge s\) in the third \("\le "\).

The Cauchy–Schwarz inequality gives us

$$\begin{aligned}&\left| \langle (\nabla _{e_\alpha }T)(du(e_\beta ), du(e_\beta )), du(e_\alpha ) \rangle \right| \le \Vert \nabla T\Vert _{L^\infty } |du|^2 \cdot |du| \\&\quad \le \varepsilon _2 e(u) + \frac{1}{4\varepsilon _2}\Vert \nabla T\Vert ^2_{L^\infty } e(u)^2, \\&\left| \langle 2T((\nabla _{e_\alpha }du)(e_\beta ), du(e_\beta )), du(e_\alpha ) \rangle \right| \le 2\Vert T\Vert _{L^\infty } |\nabla du||du| \cdot |du| \le \varepsilon _3 |\nabla du|^2 \\&\quad + \frac{1}{\varepsilon _3}\Vert T\Vert ^2_{L^\infty } e(u)^2. \end{aligned}$$

The formula (3.12) in [2] (see also [3]) implies that for any \(\epsilon >0\)

$$\begin{aligned} |\nabla du|^2 \ge \frac{1-\epsilon }{m-1}|\tau (u)|^2 + \left( \frac{m}{m-1}- \frac{1}{(m-1)\epsilon } \right) |\nabla \sqrt{e(u)}|^2. \end{aligned}$$

Choosing \(\epsilon = m\), then we have

$$\begin{aligned} |\nabla du|^2 \ge -|\tau (u)|^2 + \left( 1+ \frac{1}{m} \right) |\nabla \sqrt{e(u)}|^2. \end{aligned}$$

By the VT-harmonic map equation (1.1), it is easy to see that

$$\begin{aligned} |\tau (u)|^2 \le e(u)\Vert V\Vert ^2_{L^\infty } + e(u)^2\Vert T\Vert ^2_{L^\infty }. \end{aligned}$$

Hence, from the Bochner formula (6.1), we obtain

$$\begin{aligned} \begin{aligned} \frac{1}{2}\Delta _V e(u) \ge&\left( -(1-\varepsilon _3) \Vert V\Vert ^2_{L^\infty } -A - \varepsilon _2\right) e(u) +(1-\varepsilon _3) \left( 1+ \frac{1}{m} \right) |\nabla \sqrt{e(u)}|^2 \\&+\left( -(1-\varepsilon _3) \Vert T\Vert ^2_{L^\infty } - \kappa \left( 1- \frac{1}{s_0}\right) - \frac{1}{4\varepsilon _2}\Vert \nabla T\Vert ^2_{L^\infty } -\frac{1}{\varepsilon _3}\Vert T\Vert ^2_{L^\infty }\right) e(u)^2. \end{aligned} \end{aligned}$$

Since

$$\begin{aligned} \frac{1}{2}\Delta _V e(u) = \frac{1}{2}\Delta _V |\nabla u|^2 = |\nabla |\nabla u||^2 + |\nabla u|\Delta _V |\nabla u|, \end{aligned}$$

therefore

$$\begin{aligned} \begin{aligned} \Delta _V |\nabla u| \ge&\left[ (1-\varepsilon _3) \left( 1+ \frac{1}{m} \right) -1\right] \frac{ |\nabla |\nabla u||^2 }{|\nabla u|} + \left( -(1-\varepsilon _3) \Vert V\Vert ^2_{L^\infty } -A - \varepsilon _2\right) |\nabla u|\\&+\left[ -(1-\varepsilon _3) \Vert T\Vert ^2_{L^\infty } - \kappa \left( 1- \frac{1}{s_0}\right) - \frac{1}{4\varepsilon _2}\Vert \nabla T\Vert ^2_{L^\infty } -\frac{1}{\varepsilon _3}\Vert T\Vert ^2_{L^\infty }\right] |\nabla u|^3. \end{aligned} \end{aligned}$$

Choose \(\varepsilon _3=\frac{1}{(m+1)^2}\), then

$$\begin{aligned} \begin{aligned} \Delta _V |\nabla u| \ge&\frac{1}{m+1} \frac{ |\nabla |\nabla u||^2 }{|\nabla u|} + \left[ -\left( 1-\frac{1}{(m+1)^2}\right) \Vert V\Vert ^2_{L^\infty } -A - \varepsilon _2\right] |\nabla u|\\&+\left[ -\left( 1-\frac{1}{(m+1)^2}\right) \Vert T\Vert ^2_{L^\infty } - \kappa \left( 1- \frac{1}{s_0}\right) \right. \\&\left. - \frac{1}{4\varepsilon _2}\Vert \nabla T\Vert ^2_{L^\infty } -(m+1)^2\Vert T\Vert ^2_{L^\infty }\right] |\nabla u|^3. \end{aligned} \end{aligned}$$
(6.8)

For simplicity in the following computation, we denote

$$\begin{aligned} \begin{aligned} C_1:=&\frac{1}{m+1}, \ \ C_2:= -\left( 1-\frac{1}{(m+1)^2}\right) \Vert V\Vert ^2_{L^\infty } -A - \varepsilon _2, \\ C_3:=&-\left( 1-\frac{1}{(m+1)^2}\right) \Vert T\Vert ^2_{L^\infty } - \kappa \left( 1- \frac{1}{s_0}\right) - \frac{1}{4\varepsilon _2}\Vert \nabla T\Vert ^2_{L^\infty } -(m+1)^2\Vert T\Vert ^2_{L^\infty }. \end{aligned} \end{aligned}$$

By direct calculation, we have

$$\begin{aligned} \begin{aligned} \nabla \psi =&\frac{\nabla |\nabla u|}{\varphi \circ u} - \frac{|\nabla u|\nabla (\varphi \circ u)}{(\varphi \circ u)^2}, \\ \Delta _V \psi =&\frac{\Delta _V |\nabla u|}{\varphi \circ u} - \frac{|\nabla u|\Delta _V(\varphi \circ u)}{(\varphi \circ u)^2} - \frac{2}{\varphi \circ u}\langle \nabla \psi , \nabla (\varphi \circ u) \rangle \\ \ge&C_1 \frac{ |\nabla |\nabla u||^2 }{(\varphi \circ u)|\nabla u|} + C_2 \frac{|\nabla u|}{\varphi \circ u}+ C_3 \frac{|\nabla u|^3}{\varphi \circ u} - \frac{\psi \Delta _V (\varphi \circ u)}{\varphi \circ u} - \frac{2\langle \nabla \psi , \nabla (\varphi \circ u) \rangle }{\varphi \circ u}. \end{aligned} \end{aligned}$$

The Cauchy–Schwarz implies that

$$\begin{aligned} \begin{aligned}&- \frac{2\langle \nabla \psi , \nabla (\varphi \circ u) \rangle }{\varphi \circ u}\\&\quad = -(2-2C_1)\frac{\langle \nabla \psi , \nabla (\varphi \circ u)\rangle }{\varphi \circ u} - 2C_1 \frac{\langle \nabla \psi , \nabla (\varphi \circ u) \rangle }{\varphi \circ u} \\&\quad = -(2-2C_1)\frac{\langle \nabla \psi , \nabla (\varphi \circ u)\rangle }{\varphi \circ u} - 2C_1 \frac{\langle \nabla |\nabla u|, \nabla (\varphi \circ u) \rangle }{(\varphi \circ u)^2} + 2C_1 \frac{|\nabla (\varphi \circ u)|^2|\nabla u|}{(\varphi \circ u)^3} \\&\quad \ge -(2-2C_1)\frac{\langle \nabla \psi , \nabla (\varphi \circ u)\rangle }{\varphi \circ u} - C_1 \frac{ |\nabla |\nabla u||^2 }{(\varphi \circ u)|\nabla u|} + C_1 \frac{|\nabla (\varphi \circ u)|^2|\nabla u|}{(\varphi \circ u)^3}. \end{aligned} \end{aligned}$$

From the above two inequalities and (6.4), we get

$$\begin{aligned} \begin{aligned} \frac{\Delta _V \psi }{\psi } \ge&C_2 + C_3 |\nabla u|^2 - \frac{\Delta _V (\varphi \circ u)}{\varphi \circ u} +C_1 \frac{|\nabla (\varphi \circ u)|^2}{(\varphi \circ u)^2} - (2-2C_1)\frac{\langle \nabla r^2, \nabla (\varphi \circ u)\rangle }{(a^2-r^2)(\varphi \circ u)}. \end{aligned}\nonumber \\ \end{aligned}$$
(6.9)

Since

$$\begin{aligned} \left| \frac{\langle \nabla r^2, \nabla (\varphi \circ u)\rangle }{(a^2-r^2)(\varphi \circ u)} \right| \le \frac{2r|\nabla (\varphi \circ u)|\cdot |\nabla r|}{(a^2-r^2)(\varphi \circ u)} \le \frac{2r|\nabla (\varphi \circ u)|}{(a^2-r^2)(\varphi \circ u)} \le \frac{2\sqrt{\kappa }r|\nabla u|}{(a^2-r^2)(\varphi \circ u)}\nonumber \\ \end{aligned}$$
(6.10)

and from (6.3),

$$\begin{aligned} \begin{aligned} \Delta _V(\varphi \circ u) =&g^{\alpha \beta }\nabla ^2 \varphi (\partial _\alpha u, \partial _\beta u) + \langle (\nabla \varphi )\circ u, \tau (u) + du(V) \rangle \\ =&g^{\alpha \beta }u^j_{x^\alpha }u^k_{x^{\beta }}\mathrm{Hess}^N(\varphi )\left( \frac{\partial }{\partial y^j}, \frac{\partial }{\partial y^k} \right) + \langle (\nabla \varphi )\circ u, -\mathrm{Tr}_gT(du, du) \rangle \qquad \\ \le&\left( -\kappa \cos (\sqrt{\kappa }\rho ) + \sqrt{\kappa }\Vert T\Vert _{L^\infty } \right) |\nabla u|^2. \end{aligned} \end{aligned}$$
(6.11)

Therefore, from (6.6), (6.7), (6.9)–(6.11), we obtain

$$\begin{aligned} \begin{aligned}&C_2+\left( C_3+\kappa -\frac{\sqrt{\kappa }\Vert T\Vert _{L^\infty }}{\cos (\sqrt{\kappa }\rho \circ u)} \right) |\nabla u|^2 - \frac{4(1-C_1)\sqrt{\kappa }r}{(a^2-r^2)(\varphi \circ u)}|\nabla u|\\&\quad - \frac{2Ar^2-2Ar_0r+2{\widetilde{C}}_0 r+2}{a^2-r^2} - \frac{8r^2}{(a^2-r^2)^2} \le 0. \end{aligned} \end{aligned}$$
(6.12)

Since the condition (6.2) tells us

$$\begin{aligned} \left( 1+(m+1)^2 - \frac{1}{(m+1)^2} \right) \Vert T\Vert ^2_{L^\infty } + \frac{\sqrt{\kappa }}{\cos (\sqrt{\kappa }\widetilde{R)}}\Vert T\Vert _{L^\infty } < \frac{\kappa }{s_0}, \end{aligned}$$

there exists a constant \(\varepsilon _0>0\), such that

$$\begin{aligned} \left( 1+(m+1)^2 - \frac{1}{(m+1)^2} \right) \Vert T\Vert ^2_{L^\infty } + \frac{\sqrt{\kappa }}{\cos (\sqrt{\kappa }\widetilde{R)}}\Vert T\Vert _{L^\infty }+\varepsilon _0 < \frac{\kappa }{s_0}. \end{aligned}$$

Choosing \(\varepsilon _2= \frac{\Vert \nabla T\Vert ^2_{L^\infty }+1}{4\varepsilon _0}\), then we have

$$\begin{aligned} \begin{aligned}&C_3+\kappa -\frac{\sqrt{\kappa }\Vert T\Vert _{L^\infty }}{\cos (\sqrt{\kappa }\rho \circ u)}= -\left( 1-\frac{1}{(m+1)^2}\right) \Vert T\Vert ^2_{L^\infty } - \kappa \left( 1- \frac{1}{s_0}\right) \\&\qquad - \frac{1}{4\varepsilon _2}\Vert \nabla T\Vert ^2_{L^\infty } -(m+1)^2\Vert T\Vert ^2_{L^\infty } +\kappa -\frac{\sqrt{\kappa }\Vert T\Vert _{L^\infty }}{\cos (\sqrt{\kappa }\rho \circ u)}\\&\quad \ge \frac{\kappa }{s_0} -\left( 1-\frac{1}{(m+1)^2}\right) \Vert T\Vert ^2_{L^\infty } - \frac{1}{4\varepsilon _2}\Vert \nabla T\Vert ^2_{L^\infty } -(m+1)^2\Vert T\Vert ^2_{L^\infty } -\frac{\sqrt{\kappa }\Vert T\Vert _{L^\infty }}{\cos (\sqrt{\kappa }{\widetilde{R}})}\\&\quad = \frac{\kappa }{s_0} -\left[ \left( 1+(m+1)^2- \frac{1}{(m+1)^2} \right) \Vert T\Vert ^2_{L^\infty } + \frac{1}{4\varepsilon _2}\Vert \nabla T\Vert ^2_{L^\infty } + \frac{\sqrt{\kappa }\Vert T\Vert _{L^\infty }}{\cos (\sqrt{\kappa }{\widetilde{R}})} \right] \\&\quad> \frac{\kappa }{s_0} -\left[ \left( 1+(m+1)^2- \frac{1}{(m+1)^2} \right) \Vert T\Vert ^2_{L^\infty } + \varepsilon _0 + \frac{\sqrt{\kappa }\Vert T\Vert _{L^\infty }}{\cos (\sqrt{\kappa }{\widetilde{R}})} \right] =: C_4 >0. \end{aligned} \end{aligned}$$

Therefore, it follows from (6.12) that

$$\begin{aligned} \begin{aligned}&C_4|\nabla u|^2 - \frac{4(1-C_1)\sqrt{\kappa }r}{(a^2-r^2)(\varphi \circ u)}|\nabla u| -\left( \frac{2Ar^2-2Ar_0r+2{\widetilde{C}}_0 r+2}{a^2-r^2} + \frac{8r^2}{(a^2-r^2)^2}-C_2\right) \\&\quad \le 0. \end{aligned} \end{aligned}$$

Note the elementary fact that if \(ax^2-bx -c \le 0\) with abc all positive, then

$$\begin{aligned} x \le \frac{b}{a} + \sqrt{\frac{c}{a}}. \end{aligned}$$

Hence, at the point q,

$$\begin{aligned} \begin{aligned} |\nabla u| \le&\frac{4(1-C_1)\sqrt{\kappa }r}{C_4(a^2-r^2)\cos (\sqrt{\kappa } {\widetilde{R}})} +\sqrt{\frac{1}{C_4}\left( \frac{2Ar^2-2Ar_0r+2{\widetilde{C}}_0 r+2}{a^2-r^2} + \frac{8r^2}{(a^2-r^2)^2} -C_2 \right) }. \end{aligned} \end{aligned}$$

From this, we can derive the upper bound of f, and it is easy to conclude that at every point of \(B_{\frac{a}{2}}({\tilde{p}})\), we have

$$\begin{aligned} \begin{aligned} |\nabla u| \le&C_5 \left( \sqrt{A+\frac{\Vert \nabla T\Vert ^2_{L^\infty }+1}{4\varepsilon _0}+ \left( 1- \frac{1}{(m+1)^2}\right) \Vert V\Vert _{L^\infty }} + \frac{1}{a} +\frac{1}{\sqrt{a}} \right) \\ \le&C_6 \left( \sqrt{A} + 1 + \frac{1}{a}+ \frac{1}{\sqrt{a}} \right) , \end{aligned} \end{aligned}$$
(6.13)

where \(C_6>0\) is a constant depending only on \(m, n, \kappa , {\widetilde{R}}, V, T\).

For any fixed \(x\in M\), letting \(a\rightarrow \infty \) in (6.13), we obtain \(|\nabla u| \le C_6(\sqrt{A} + 1 )\). \(\square \)

Proof of Theorem 4

We first choose a constant \(C'_0\) depending only on \(\kappa , \sigma , R\) and the geometry of N, such that if

$$\begin{aligned} \max |\nabla T| + \max |T| \le C'_0, \end{aligned}$$

then both the condition (4.3) in Theorem 5 and the condition (6.2) in Theorem 9 hold. Let \(\{\Omega _i\}\) be a compact exhaustion of M. By Theorem 5, we have a sequence of maps \(\{u_i\}\) which solve the Dirichlet problem

$$\begin{aligned} \left\{ \begin{array}{rll} \displaystyle &{} \tau (u_i) + du_i(V) + \mathrm{Tr}_g T(du_i, du_i) = 0, \\ \displaystyle &{} \left. u_i\right| _{\partial \Omega _i}= \left. u_0\right| _{\partial \Omega _i}, \\ &{} u_i \ \ \mathrm{homotopic} \ \ \mathrm{to} \ \ u_0 \ \ \mathrm{rel.} \ \ \partial \Omega _i, \end{array} \right. \end{aligned}$$

where \(u_i \in C^\infty (\Omega _i, N)\cap C^0(\overline{\Omega _i}, N)\) such that \(u_i(\Omega _i)\subset B_R(p)\).

For any compact set \(K \subset M\), there exists an integer \(i_0 >0\), such that \(K\subset \Omega _i\) for \(i>i_0\). Then, by (6.8),

$$\begin{aligned} \begin{aligned} \Delta _V |\nabla u_i| \ge&\frac{1}{m+1} \frac{ |\nabla |\nabla u_i||^2 }{|\nabla u_i|} + \left[ -\left( 1-\frac{1}{(m+1)^2}\right) \Vert V\Vert ^2_{L^\infty } -{\widetilde{A}} - \frac{\Vert \nabla T\Vert ^2_{L^\infty }+1}{4\varepsilon _0}\right] |\nabla u_i|\\&+\left[ -\left( 1+(m+1)^2-\frac{1}{(m+1)^2}\right) \Vert T\Vert ^2_{L^\infty } - \kappa \left( 1- \frac{1}{s_0}\right) - \varepsilon _0 \right] |\nabla u_i|^3, \end{aligned}\nonumber \\ \end{aligned}$$
(6.14)

where \({\widetilde{A}}\) is a positive constant depending only on the bounds for Ricci curvature of K and \(\Vert V\Vert _{C^1(K)}\).

Since K is compact, there exist finitely many such geodesic balls \(\{B_{a_j}({\tilde{p}}_j)\}_{j=1}^{k_0}\subset M\), such that \(\cup _{j=1}^{k_0}B_{a_j}({\tilde{p}}_j)\supset K\). Hence, for any \(q\in K\), there is a geodesic ball, say \(B_{a_{j_0}}({\tilde{p}}_{j_0}) \ \ (1\le j_0 \le k_0)\), containing q. Then, by Theorem 9, we can conclude that

$$\begin{aligned} |\nabla u_i|(q) \le C_{7j_0}(\sqrt{{\widetilde{A}}}+1). \end{aligned}$$

Hence,

$$\begin{aligned} \sup _{K}|\nabla u_i| \le \max _{1\le j_0 \le k_0} \left\{ C_{7j_0}(\sqrt{{\widetilde{A}}}+1) \right\} =: C_8, \end{aligned}$$

where \(C_8\) is a positive constant independent of i. Then, by the standard elliptic theory, \(u_i\) subconverges to a VT-harmonic map \(u \in C^\infty (M, N)\) with \(u(M)\subset B_R(p)\) and u is homotopic to \(u_0\). \(\square \)