1 Introduction

In 1976, Nakagawa and Takagi studied some properties about Kähler imbeddings of compact Hermitian symmetric spaces in the complex projective space \(CP^n\) and gave a classification of Kähler submanifolds in \(CP^n\) with parallel second fundamental form (cf. [8]). In 1984, Ros decided the compact Einstein Kähler submanifold in \(CP^n\) with parallel second fundamental form (cf. [9]). In 1985, Tsukada classified \(2n\)-dimensional totally complex submanifolds in \(HP^n\) with parallel second fundamental form into eight types (cf. [10, 11]). Recently, we studied conformal minimal immersions of 2-spheres in \(CP^n\) and \(G(k,N)\) with parallel second fundamental form, and obtained some geometric properties of them (cf. [6, 7]).

In this paper, our interest is to study classification of conformal minimal immersions from \(S^2\) to the quaternionic projective space \(HP^n\) with parallel second fundamental form by the theory of harmonic maps. Here, we regard \(HP^n\) as a totally geodesic totally real submanifolds in complex Grassmann manifolds \(G(2,2n+2)\) and obtain the following classification theorem:

Theorem 1.1

Let \(\varphi :S^2\rightarrow HP^n\) be a linearly full conformal minimal immersion, and let \(K\) and \(B\) be its Gauss curvature and second fundamental form respectively. If \(B\) is parallel, then, up to a symplectic isometry of \(HP^n\), it belongs to one of the following minimal immersions.

  1. (1)

    \(\varphi =\left[ (\sqrt{3}\overline{z}+\sqrt{3}\overline{z}^2j,~-1+ 2z\overline{z}-2\overline{z}j-z\overline{z}^2j)^T\right] : S^2\rightarrow HP^1\), with \(K=\frac{2}{3},\Vert B\Vert ^2=\frac{8}{3}\);

  2. (2)

    \(\varphi =\left[ (1,~z)^T\right] :S^2\rightarrow CP^1\subset HP^1\), with \(K=2, ~B=0\);

  3. (3)

    \(\varphi =\left[ (1,~\sqrt{2}z,z^2)^T\right] :S^2\rightarrow CP^2\subset HP^2\), with \(K=1, ~\Vert B\Vert ^2=2\);

  4. (4)

    \(\varphi =\left[ (1-\frac{1}{2}\overline{z}^3j, ~\sqrt{3}z+ \frac{\sqrt{3}}{2}\overline{z}^2j, ~\frac{3}{2}z^2,~\frac{\sqrt{3}}{2}z^3)^T\right] : S^2\rightarrow HP^3\), with \(K=\frac{2}{3}, ~\Vert B\Vert ^2=\frac{1}{3}\);

  5. (5)

    \(\varphi =\left[ (-2\overline{z},~\sqrt{2}-\sqrt{2}z\overline{z},~2z)^T\right] :S^2\rightarrow CP^2\subset HP^2\), with \(K=\frac{1}{2}, ~B=0\);

  6. (6)

    \(\varphi =\left[ (6\overline{z}^2,~-6\overline{z}+6z\overline{z}^2, ~\sqrt{6}-4\sqrt{6}z\overline{z}+ \sqrt{6}z^2\overline{z}^2,~6z-6z^2\overline{z}, ~6z^2)^T\right] :S^2\rightarrow CP^4\subset HP^4\), with \(K=\frac{1}{6}, ~\Vert B\Vert ^2 = \frac{2}{3}\).

Further, no two of the above six cases are congruent, i.e., there is no symplectic isometry which transforms one case into another.

2 Preliminaries

(A) For any \(N=1,2,\ldots ,\) let \(\left\langle ,\right\rangle \) denote the standard Hermitian inner product on \(\mathbb {C}^{N}\) defined by \(\left\langle z, w\right\rangle = z_1\overline{w}_1 +\cdots + z_N\overline{w}_N\), where \(z=(z_1, \ldots , z_N)^T, w=(w_1, \ldots , w_N)^T\in {\mathbb {C}}^N\) and \(\bar{\,}\) denote complex conjugation. Let \(\mathbb {H}\) denote the division ring of quaternions. Let \(j\) be a unit quaternion with \(j^2=-1\). Then, we have an identification of \({\mathbb {C}}^2\) with \(\mathbb {H}\) given by making \((a,b)\in {\mathbb {C}}^2\) correspond to \(a+bj\in \mathbb {H}\); let \(n \in \left\{ 1,2,\ldots \right\} \), and we have a corresponding identification of \({\mathbb {C}}^{2n}\) with \({\mathbb {H}}^n\). For any \(a+bj\in \mathbb {H}\), the left multiplication by \(j\) is given by \(j(a+bj)=-\overline{b}+\overline{a}j\); the conjugation is given by \(\overline{a+bj}=\overline{a}-bj\); the positive-definite inner product is given by \(\left\langle x,y\right\rangle _{\mathbb {H}}=Re(x\overline{y})\) for any \(x,y\in \mathbb {H}\).

Let \(\mathbf {J}:{\mathbb {C}}^{2n} \rightarrow {\mathbb {C}}^{2n}\) be the conjugate linear map given by left multiplication by \(j\), i.e.,

$$\begin{aligned} {\mathbf {J}}(z_1, z_2, \ldots , z_{2n-1}, z_{2n})^T= (-\bar{z}_2, \bar{z}_1, \ldots , -\bar{z}_{2n}, \bar{z}_{2n-1})^T. \end{aligned}$$

Then, \({\mathbf {J}}^2=-id\) where \(id\) denotes the identity map on \({\mathbb {C}}^{2n}\). In fact, for any \(v\in {\mathbb {C}}^{2n}\),

$$\begin{aligned} {\mathbf {J}}v=J_n \bar{v}, \end{aligned}$$

where \(J_n=diag \underbrace{\left\{ \left( \begin{array} {lr} 0 &{} -1 \\ 1 &{} 0 \\ \end{array}\right) , \ldots , \left( \begin{array} {lr} 0 &{} -1 \\ 1 &{} 0 \\ \end{array}\right) \right\} }_{n}\).

By the above, we immediately have the following lemma (cf. [1]).

Lemma 2.1

The operator \(\mathbf {J}\) has the following properties:

  1. (i)

    \(\left\langle {\mathbf {J}}v, {\mathbf {J}}w\right\rangle =\left\langle w, v\right\rangle \) for all \(v,w\in {\mathbb {C}}^{2n}\);

  2. (ii)

    \(\left\langle {\mathbf {J}}v, v\right\rangle =0\) for all \(v\in {\mathbb {C}}^{2n}\);

  3. (iii)

    \(\partial \circ {\mathbf {J}}=\mathbf {J} \circ \overline{\partial }\), \(\overline{\partial } \circ \mathbf {J}=\mathbf {J} \circ \partial \);

  4. (iv)

    \(\mathbf {J}(\lambda v)=\overline{\lambda } \mathbf {J} v\) for any \(\lambda \in {\mathbb {C}}\), \(v\in {\mathbb {C}}^{2n}\).

Let \(G(2,2n+2)\) denote the Grassmann manifold of all complex \(2\)-dimensional subspaces of \({\mathbb {C}}^{2n+2}\) with its standard Kähler structure. The quaternionic projective space \(HP^{n}\) is the set of all one-dimensional quaternionic subspaces of \({\mathbb {H}}^{n+1}\). Throughout the above, we shall regard \(HP^{n}\) as the totally geodesic submanifold of \(G(2,2n+2)\) given by

$$\begin{aligned} HP^{n}=\left\{ V\in {G(2,2n+2)}: {\mathbf {J}}V=V\right\} . \end{aligned}$$

Let \(Sp(n+1)=\left\{ g\in GL(n+1;\mathbb {H}), \ g^*g=I_{n+1}\right\} \) be the symplectic isometry group of \(HP^{n}\), here \(I_{n+1}\) is the identity matrix. The explicit description is that the following diagram commutes:

where \(i_1,i_2\) are inclusions and \(\pi _1,\pi _2\) are projections, and \(i_1(g)=U\), for \(1\le a,b\le n+1\)

$$\begin{aligned} {\left\{ \begin{array}{ll} U_{2b-1}^{2a-1}=A_{b}^{a},\quad U_{2b}^{2a-1}=-\overline{D}_{b}^{a},\\ U_{2b-1}^{2a}=D_{b}^{a},\quad U_{2b}^{2a}=\overline{A}_{b}^{a}, \end{array}\right. } \end{aligned}$$

where \(A=(A_{b}^{a}),~D=(D_{b}^{a})\in M_{n+1}(\mathbb {C}),~g=A+\overline{D}j\in Sp(n+1);\)

Here, we take the standard metric on \(G(2,2n+2)\) as described in section 2 of [7]; then, the metric induced by \(i_2\) is twice as much as the standard metric on \(HP^{n}\).

Thus, a harmonic map from \(S^2\) to \(HP^n\) is precisely a harmonic map from \(S^2\) to \(G(2,2n+2)\) which has image in \(i_2(HP^n)\).

Then, for any \(g\in Sp(n+1)\), the action of \(g\) on \(HP^{n}\) induces an action of \(i_1(g)\) on \(CP^{2n+1}\), where \(i_1(g)\in U(2n+2)\) commutes with \({\mathbf {J}}\). We shall retain \(g\) to also denote \(i_1(g)\). Then

$$\begin{aligned} Sp(n+1)=\left\{ g\in U(2n+2), \ g\circ {\mathbf {J}}={\mathbf {J}} \circ g\right\} =\left\{ g\in U(2n+2), \ g J_{n+1} g^{T}=J_{n+1}\right\} . \end{aligned}$$

In the following, we deal with the symplectic isometry of \(HP^{n}\) through the corresponding symplectic isometry of \(CP^{2n+1}\).

(B) In this section, we give general expression of some geometric quantities about conformal minimal immersions from \(S^2\) to \(HP^n\) (cf. [7]).

Let \(M\) be a simply connected domain in the unit sphere \(S^2\) and let \((z,\overline{z})\) be complex coordinates on \(M\). We take the metric \(ds^2_M=dzd\overline{z}\) on \(M\). Denote

$$\begin{aligned} \partial =\frac{\partial }{\partial z}, \quad \overline{\partial }=\frac{\partial }{\partial \overline{z}}. \end{aligned}$$

We consider the complex Grassmann manifold \(G(2,N)\) as the set of Hermitian orthogonal projections from \({\mathbb {C}}^N\) onto a \(2\)-dimensional subspace in \({\mathbb {C}}^N\). Then, a map \(\psi :S^2\rightarrow G(2,N)\) is a Hermitian orthogonal projection onto a \(2\)-dimensional subbundle \(\underline{\psi }\) of the trivial bundle \({\underline{\mathbb {C}}}^N=M\times {\mathbb {C}}^N\) given by setting the fibre of \(\underline{\psi }\) at \(x\), \({\underline{\psi }}_x\), equal to \(\psi (x)\) for all \(x\in M\). We say that \(\underline{\psi }\) is a harmonic subbundle if \(\psi \) is harmonic (cf. [3]).

Let \(\varphi : S^2 \rightarrow HP^{n}\) be a conformal minimal immersion. The map \(i_2\circ \varphi :S^2 \rightarrow G(2,2n+2)\) may be represented via the local sections of the subbundle \(\underline{Im}(i_2\circ \varphi )\) by the projection map (cf. [7], (2.10)):

$$\begin{aligned} i_2\circ \varphi =XX^*+({\mathbf {J}}X)({\mathbf {J}}X)^*, \end{aligned}$$

where \(X\in \underline{Im}(i_2\circ \varphi )\) is a unit column vector in \(\mathbb {C}^{2n+2}\), and \(X\) and \({\mathbf {J}}X\) are naturally orthogonal.

Denote \(i_2\circ \varphi \) by \(\varphi _0\) (in the following, we will use this notation all the time). Suppose that the metric induced by \(\varphi _0\) is \(ds^2=\lambda ^2dzd\overline{z}\). Let \(K\) and \(B\) be its Gauss curvature and second fundamental form, respectively. From section 2 and 3 of [7], we have

$$\begin{aligned} {\left\{ \begin{array}{ll} \lambda ^2=tr\partial {\varphi _0}\overline{\partial }{\varphi _0},\\ K=-\frac{2}{\lambda ^2} \partial \overline{\partial }\log \lambda ^2,\\ \Vert B \Vert ^2=4trPP^*. \end{array}\right. } \end{aligned}$$
(2.1)

where \(P=\partial \left( {A_z}/{\lambda ^2}\right) \) with \(A_z=\left( 2\varphi _0-I\right) \partial {\varphi _0}\), and \(I\) is the identity matrix, then \(P^*=\overline{\partial }\left( {A_z^*}/{\lambda ^2}\right) ,~A_z^*=-A_{\overline{z}}\).

3 The proof of main theorem

We recall that an immersion of \(S^2\) in \(HP^n\) is conformal and minimal if and only if it is harmonic (cf. [4], Sec 10.6). Thus, we shall consider the immersive harmonic maps from \(S^2\) to \(HP^n\) with parallel second fundamental form for the reducible and irreducible cases to give the proof of Theorem 1.1 in Sect. 1. At first, we state a conclusion about parallel minimal immersions of 2-spheres in \(G(k,N)\) as follows:

Lemma 3.1

([7]) Let \(\varphi :S^2 \rightarrow G(k,N)\) be a conformal minimal immersion with the second fundamental form \(B\). Then \(B\) is parallel if and only if the equation

$$\begin{aligned} \frac{\lambda ^2}{16}\Vert B \Vert ^2\left( 8K+\Vert B \Vert ^2\right) +2tr[A_z, P][A_z^*, P^*]-5tr[A_z, A_z^*][P, P^*]=0 \end{aligned}$$
(3.1)

holds.

(I) Let \(\varphi :S^2\rightarrow HP^n\) be a linearly full reducible harmonic map, then by ([1], Proposition 3.7) we know that \(\varphi \) is a quaternionic mixed pair or a quaternionic Frenet pair. In the following, we discuss the two cases of \(\varphi \) with parallel second fundamental form, respectively.

(Ia) If \(\varphi \) is a linearly full quaternionic Frenet pair, then

$$\begin{aligned} \underline{\varphi }_0=\underline{f}_n^{(2n+1)} \oplus \underline{f}_{n+1}^{(2n+1)}, \end{aligned}$$
(3.2)

where \(\underline{f}_0^{(2n+1)},~\underline{f}_1^{(2n+1)}, \ldots , ~\underline{f}_{2n+1}^{(2n+1)}:S^2 \rightarrow CP^{2n+1}\) is the harmonic sequence generated by a linearly full totally \(\mathbf {J}\)-isotropic map \(\underline{f}_0^{(2n+1)}\).

Firstly, we recall ([1], §3) that a full holomorphic map \(\underline{f}_0^{(2n+1)} : S^2 \rightarrow CP^{2n+1}\) in the following harmonic sequence satisfying \({\underline{f}}_{2n+1}^{(2n+1)}={\mathbf {J}}\underline{f}_0^{(2n+1)}\) is said to be totally \(\mathbf {J}\)-isotropic,

$$\begin{aligned} 0\mathop {\longleftarrow }\limits ^{A_0''} {\underline{f}} _0^{(2n+1)} \mathop {\longrightarrow }\limits ^{A_0'}\cdots \mathop {\longrightarrow }\limits ^{A_{n-1}'} {\underline{f}}_{n}^{(2n+1)} \mathop {\longrightarrow }\limits ^{A_{n}'} {\underline{f}}_{n+1}^{(2n+1)} \mathop {\longrightarrow }\limits ^{A_{n+1}'} \cdots \mathop {\longrightarrow }\limits ^{A_{2n}'} {\underline{f}}_{2n+1}^{(2n+1)} \mathop {\longrightarrow }\limits ^{A_{2n+1}'} 0, \end{aligned}$$

where \(A'_{j}(v)=\pi _{{f_{j}^{(2n+1)}}^{\perp }}(\partial v), \quad A''_{j}(v)=\pi _{{f_{j}^{(2n+1)}}^{\perp }}(\overline{\partial } v)\) for \(v\in C^{\infty }({\underline{f}}_{j}^{(2n+1)})\), here \(\pi _{{f_{j}^{(2n+1)}}^{\perp }}\) denotes orthogonal projection onto bundle \({{\underline{f}}_{j}^{(2n+1)}}^{\perp }\) and \(C^{\infty }({\underline{f}}_{j}^{(2n+1)})\) denotes the vector space of smooth sections of bundle \({\underline{f}}_{j}^{(2n+1)}\), \(j=0,\ldots ,2n+1\).

Let \(f_0^{(2n+1)}\) be a holomorphic section of \({\underline{f}} _0^{(2n+1)}\), i.e., \(\overline{\partial } f_0 ^{(2n+1)} = 0\), and let \(f_j^{(2n+1)}\) be a local section of \({\underline{f}}_{j}^{(2n+1)}\) such that

$$\begin{aligned} f_j^{(2n+1)} = \pi _{{f_{j-1}^{(2n+1)}}^{\perp }}\left( \partial f_{j-1} ^{(2n+1)} \right) \end{aligned}$$

for \(j=1, \ldots ,2n+1\). Then, we have some formulas as follows (cf. [2]):

$$\begin{aligned} \partial f_j ^{(2n+1)}&= f_{j+1}^{(2n+1)} + \partial \log |f_j ^{(2n+1)}|^2 f_j ^{(2n+1)}, \ j= 0,\ldots , 2n,\\ \overline{\partial } f_j^{(2n+1)}&= - l_{j-1} ^{(2n+1)} f_{j-1} ^{(2n+1)}, \ j = 1, \ldots , 2n+1,\\ \partial \overline{\partial } \log |f_j ^{(2n+1)}|^2&= l_j^{(2n+1)}-l_{j-1}^{(2n+1)},\\ \partial \overline{\partial } \log l_j^{(2n+1)}&= l_{j+1} ^{(2n+1)}- 2 l_j^{(2n+1)} + l_{j-1}^{(2n+1)}, \ j =0, \ldots , 2n, \end{aligned}$$

where \(l_j ^{(2n+1)}= |f_{j+1}^{(2n+1)}|^2/ |f_j^{(2n+1)}|^2\) for \(j =0, \ldots , 2n+1\), and \(l_{-1} ^{(2n+1)}= l_{2n+1}^{(2n+1)}=0\).

Since \({\underline{f}} _0^{(2n+1)}\) is totally \(\mathbf {J}\)-isotropic, in a similar fashion to ([2], Lemma 7.1) we obtain

$$\begin{aligned} l_j^{(2n+1)}=l_{2n-j}^{(2n+1)}. \end{aligned}$$
(3.3)

And set \(\mathbf {J} f_0^{(2n+1)}=\tau _0 f_{2n+1}^{(2n+1)}\) , then

$$\begin{aligned} |\tau _0|^2=\frac{|f_0^{(2n+1)}|^2}{|f_{2n+1}^{(2n+1)}|^2},\quad \mathbf {J} f_j^{(2n+1)}=(-1)^j \tau _0 \frac{|f_{2n+1}^{(2n+1)}|^2}{|f_{2n+1-j}^{(2n+1)}|^2} f_{2n+1-j}^{(2n+1)}, \end{aligned}$$

where \(j=0, \ldots , n\).

Obviously, \(\varphi _0\) belongs to the following harmonic sequence (cf. [3])

$$\begin{aligned} 0 \mathop {\longleftarrow }\limits ^{A_{0}''}\underline{f}_0^{(2n+1)} \mathop {\longleftarrow }\limits ^{A_{1}''}\cdots \mathop {\longleftarrow }\limits ^{A_{n-1}''}\underline{f}_{n-1}^{(2n+1)} \mathop {\longleftarrow }\limits ^{A_{\varphi _0}''}\underline{\varphi }_0 \mathop {\longrightarrow }\limits ^{A_{\varphi _0}'}\underline{f}_{n+2}^{(2n+1)} \mathop {\longrightarrow }\limits ^{A_{n+2}'}\cdots \mathop {\longrightarrow }\limits ^{A_{2n}'}\underline{f}_{2n+1}^{(2n+1)} \mathop {\longrightarrow }\limits ^{A_{2n+1}'}0, \end{aligned}$$
(3.4)

where \(A'_{\varphi _0}(v)=\pi _{\varphi _0^{\perp }}(\partial v), \quad A''_{\varphi _0}(v)=\pi _{\varphi _0^{\perp }}(\overline{\partial } v)\) for \(v\in C^{\infty }(\underline{\varphi }_0)\), here \(\pi _{\varphi _0^{\perp }}\) denotes orthogonal projection onto bundle \(\underline{\varphi }_0^{\perp }\) and \(C^{\infty }(\underline{\varphi }_0)\) denotes the vector space of smooth sections of bundle \(\underline{\varphi }_0\).

From (3.2), we have \(\varphi _0=\frac{{f}_n^{(2n+1)}({f}_n^{(2n+1)})^*}{|{f}_n^{(2n+1)}|^2}+\frac{{f}_{n+1}^{(2n+1)}({f}_{n+1}^{(2n+1)})^*}{|{f}_{n+1}^{(2n+1)}|^2}\). Then by (2.1), (3.3) and a series of calculations, we obtain

$$\begin{aligned} {\left\{ \begin{array}{ll} \lambda ^2=2l_{n-1}^{(2n+1)},\\ K=2-\frac{l_{n}^{(2n+1)}+l_{n-2}^{(2n+1)}}{l_{n-1}^{(2n+1)}},\\ \Vert B \Vert ^2=2\frac{l_{n}^{(2n+1)}+l_{n-2}^{(2n+1)}}{l_{n-1}^{(2n+1)}},\\ tr[A_z, P][A_z^*, P^*]=-l_{n}^{(2n+1)},\\ tr[A_z, A_z^*][P, P^*]=\frac{1}{2}l_{n-2}^{(2n+1)}. \end{array}\right. } \end{aligned}$$
(3.5)

Now, we prove that if \(\varphi :S^2 \rightarrow HP^n\) is a linearly full quaternionic Frenet pair with parallel second fundamental form, then, up to \(Sp(n+1)\), it belongs to the following case: (1) \(\varphi =\left[ (\sqrt{3}\overline{z}+\sqrt{3}\overline{z}^2j,-1+2z\overline{z}-2\overline{z}j-z\overline{z}^2j)^T\right] :S^2\rightarrow HP^1,\) with \(K=\frac{2}{3},||B||^2=\frac{8}{3}\).

If \(\varphi \) is a linearly full quaternionic Frenet pair with parallel second fundamental form, then applying Lemma 3.1 and substituting (3.5) into (3.1), we get

$$\begin{aligned} 3l_{n-2}^{(2n+1)}l_{n-1}^{(2n+1)}+4l_{n-1}^{(2n+1)}l_{n}^{(2n+1)}-3 \left( l_{n}^{(2n+1)}+l_{n-2}^{(2n+1)}\right) ^2=0. \end{aligned}$$
(3.6)

Since the second fundamental form of the map \(\varphi \) is parallel, its Gauss curvature is a constant (cf. [7], Theorem 4.5). We know up to \(U(2n+2)\), \(\underline{f}_0^{(2n+1)}\) is a Veronese surface by ([5], Lemma 4.1). Then from [2], we have \(\underline{f}_0^{(2n+1)}, \underline{f}_1^{(2n+1)}, \ldots , \underline{f}_{2n+1}^{(2n+1)}\) is the Veronese sequence in \(CP^{2n+1}\), up to \(U(2n+2)\). So, from ([2], Section 5), we get

$$\begin{aligned} |f_i^{(2n+1)}|^2=\frac{(2n+1)!i!}{(2n+1-i)!}(1+z\overline{z})^{2n+1-2i}, ~l_j ^{(2n+1)}=\frac{(j+1)(2n+1-j)}{(1+z\overline{z})^2}, \end{aligned}$$
(3.7)

where \(i=0,\ldots ,2n+1,~j=0,\ldots ,2n\).

Substituting (3.7) into (3.6), we get

$$\begin{aligned} (n-1)(n+3)(5n^2+10n-4)=0, \end{aligned}$$

which implies \(n=1\), since \(n\) is a positive integer. Hence,

$$\begin{aligned} \underline{\varphi }_0=U\underline{V}_1^{(3)} \oplus {\mathbf {J}} U\underline{V}_1^{(3)}, \end{aligned}$$
(3.8)

where \(\underline{V}_1^{(3)}\) is a Veronese surface in \(CP^3\subset CP^{2n+1}\) with the standard expression given in ([2], §5), and \(U\in U(2n+2)\) satisfies \(J_{n+1} \overline{U} \overline{V}_0^{(3)}=\lambda U V_3^{(3)}\) (\(\lambda \) is a parameter).

Set \(U^{T}J_{n+1}U=\overline{W}\), then we immediately get

$$\begin{aligned} \overline{W}V_0^{(3)}=\overline{\lambda }\overline{V}_3^{(3)},~W^{T}=-W,~W^*W=I, \end{aligned}$$
(3.9)

where \(I\) is the identity matrix.

Define a set

$$\begin{aligned} G_W\triangleq \left\{ U\in U(2n+2), \ UWU^{T}=J_{n+1}\right\} . \end{aligned}$$

For a given \(W\), the following can be easily checked

  1. (i)

    \(\forall \ g\in Sp(n+1), \ U\in G_W\), we have that \(gU\in G_W\);

  2. (ii)

    \(\forall \ U, V \in G_W, \ \exists \ g=UV^* \in Sp(n+1) \ s.t. \ U=gV\).

Then, we discuss the type of \(W\) to get the type of the corresponding \(U\). From ([2], section 5), we get

$$\begin{aligned} V^{(3)}_0&= (1, \ \sqrt{3} z,\ \sqrt{3} z^2,\ z^3,\ 0,\ldots ,\ 0)^T,\nonumber \\ V^{(3)}_3&= \frac{6}{(1+z\overline{z})^3}(-\overline{z}^3, \ \sqrt{3}\overline{z}^2, \ -\sqrt{3}\overline{z},\ 1,\ 0,\ldots ,\ 0)^T. \end{aligned}$$
(3.10)

Then, by (3.9) and (3.10) we get the type of \(\overline{W}\) as follows:

(3.11)

From \(\overline{W}\overline{U}^T = U^T J_{n+1}\), the corresponding \(U=\left[ e_1,e_2,\ldots ,e_{2n+1},e_{2n+2}\right] ^T\) satisfy

$$\begin{aligned} e_{2p}=\overline{W}\overline{e}_{2p-1},\quad p=1,\ldots ,n+1, \end{aligned}$$
(3.12)

where \(e_i\) are unit column vectors in \(\mathbb {C}^{2n+2}\).

Without loss of generality, in this case we choose

$$\begin{aligned} {\left\{ \begin{array}{ll} e_1=(1,0,0,0,\ldots ,0)^T,\\ e_3=(0,1,0,0,\ldots ,0)^T. \end{array}\right. } \end{aligned}$$
(3.13)

By(3.11)–(3.13), we get \(e_2=\overline{W}\overline{e}_1=(0,0,0,-1,\ldots ,0)^T\) and \(e_4=\overline{W}\overline{e}_3=(0,0,1,0,\ldots ,0)^T\), obviously \(\{e_1,e_2,e_3,e_4\}\) are mutually orthogonal. Next, we choose a unit column vector \(e_5=(0,0,0,0,*)^T\in \mathbb {C}^{2n+2}\), which satisfies \(\{e_1,e_2,e_3,e_4,e_5\}\) are mutually orthogonal. Set \(e_6=\overline{W}\overline{e}_5\), then \(\{e_1,e_2,e_3,e_4,e_6\}\) are mutually orthogonal. Since \(\left\langle e_6,~e_5\right\rangle =e_5^TW^Te_5=-tr(e_5e_5^TW)=0\), then \(\{e_1,e_2,e_3,e_4,e_5,e_6\}\) are mutually orthogonal.

Generally, suppose \(\{e_1,e_2,\ldots ,e_{2p-3},e_{2p-2}=\overline{W}\overline{e}_{2p-3}\}~(p\ge 3)\) are mutually orthogonal, we choose a unit column vector \(e_{2p-1}=(0,0,0,0,*)^T\in \mathbb {C}^{2n+2}\) such that \(\{e_1,e_2,\ldots ,e_{2p-3},e_{2p-2},e_{2p-1}\}\) are mutually orthogonal. Set \(e_{2p}=\overline{W}\overline{e}_{2p-1}\), then

$$\begin{aligned} \left\langle e_{2p},~e_{2p-1}\right\rangle =e_{2p-1}^TW^Te_{2p-1}=-tr(e_{2p-1}e_{2p-1}^TW)=0, \end{aligned}$$

and for any \(2\le q\le p\),

$$\begin{aligned}&\left\langle e_{2p},~e_{2q-3}\right\rangle ={e}_{2p-1}^TW^Te_{2q-3} =-{e}_{2p-1}^TWe_{2q-3} =-{e}_{2p-1}^T\overline{e}_{2q-2}\\&\quad =-\left\langle e_{2p-1},~e_{2q-2}\right\rangle =0,\left\langle e_{2p},~e_{2q-2}\right\rangle \\&\quad ={e}_{2p-1}^TW^T\overline{W}\overline{e}_{2q-3} ={e}_{2p-1}^T\overline{e}_{2q-3} =\left\langle e_{2p-1},~e_{2q-3}\right\rangle =0. \end{aligned}$$

Thus \(\{e_1,e_2,\ldots ,e_{2p-3},e_{2p-2},e_{2p-1},e_{2p}\}\) are mutually orthogonal.

So, we can choose \(n-1\) proper unit column vectors \(e_{2p+1}=(0,0,0,0,*)^T\in \mathbb {C}^{2n+2}~(2\le p\le n)\) such that \(\{e_1,e_2,\ldots ,e_{2n+1},e_{2n+2}=\overline{W}\overline{e}_{2n+1}\}\) are mutually orthogonal, and the type of the corresponding \(U\) is as follows:

(3.14)

Thus, we have

$$\begin{aligned}&UV_1^{(3)}=\frac{-\sqrt{3}}{1+z\overline{z}}(\sqrt{3}\overline{z}, \ \sqrt{3}z^2, \ -1+2z\overline{z},\ -2z+z^2\overline{z},\ 0,\ldots ,\ 0)^T,\\&{\mathbf {J}}UV_1^{(3)}=\frac{-\sqrt{3}}{1+z\overline{z}}(-\sqrt{3}\overline{z}^2, \ \sqrt{3}z,\ 2\overline{z}-z\overline{z}^2, \ -1+2z\overline{z},\ 0,\ldots ,\ 0)^T. \end{aligned}$$

Obviously, in this case \(\varphi \) is congruent to the case (1) with \(K=\frac{2}{3},~\Vert B \Vert ^2=\frac{8}{3}\).

(Ib) If \(\varphi \) is a linearly full quaternionic mixed pair, then

$$\begin{aligned} \underline{\varphi }_0=\underline{f}_0^{(m)} \oplus {\mathbf {J}} \underline{f}_0^{(m)}, \end{aligned}$$
(3.15)

where \(\underline{f}_0^{(m)} :S^2 \rightarrow CP^{m} \subseteq CP^{2n+1}\) (\(n\le m\le 2n+1\)) is holomorphic and \({\underline{f}}_1^{(m)} \perp {\mathbf {J}}{\underline{f}_0^{(m)}}\).

Obviously, \(\varphi _0\) belongs to the following harmonic sequence

$$\begin{aligned} 0\mathop {\longleftarrow }\limits ^{A_{m}''} \mathbf {J}{{\underline{f}}_m^{(m)}} \mathop {\longleftarrow }\limits ^{A_{m-1}''} \ldots \mathop {\longleftarrow }\limits ^{A_{1}''} \mathbf {J}{{\underline{f}}_1^{(m)}} \mathop {\longleftarrow }\limits ^{A_{\varphi _0}''} \underline{\varphi }_0 \mathop {\longrightarrow }\limits ^{A_{\varphi _0}'} {\underline{f}}_1^{(m)} \mathop {\longrightarrow }\limits ^{A_{1}'} \ldots \mathop {\longrightarrow }\limits ^{A_{m-1}'} {\underline{f}}_m^{(m)} \mathop {\longrightarrow }\limits ^{A_{m}'} 0. \end{aligned}$$
(3.16)

As in the case (Ia), let \(f_0^{(m)}\) be a holomorphic section of \({\underline{f}} _0^{(m)}\), i.e., \(\overline{\partial } f_0 ^{(m)} = 0\), and \(f_j^{(m)}\) \(~(j=1,\ldots ,m)\) satisfy the corresponding formulas. From (3.15), we have \(\varphi _0=\frac{{f}_0^{(m)}({f}_0^{(m)})^*}{|{f}_0^{(m)}|^2}+\frac{({\mathbf {J}} {f}_0^{(m)})({\mathbf {J}} {f}_0^{(m)})^*}{|{f}_{0}^{(m)}|^2}\). Then by (2.1) and a series of calculations, we obtain

$$\begin{aligned} {\left\{ \begin{array}{ll} \lambda ^2=2l_{0}^{(m)},\\ K=2-\frac{l_{1}^{(m)}}{l_{0}^{(m)}},\\ \Vert B \Vert ^2=2\frac{l_{1}^{(m)}}{l_{0}^{(m)}},\\ tr[A_z, P][A_z^*, P^*]=-\frac{1}{4}\frac{\big |\left\langle f_2^{(m)},~\mathbf {J}{f_1^{(m)}}\right\rangle \big |^2}{|f_1^{(m)}|^4},\\ tr[A_z, A_z^*][P, P^*]=\frac{1}{2}\left( l_{1}^{(m)}-\frac{\big |\left\langle f_2^{(m)},~\mathbf {J}{f_1^{(m)}}\right\rangle \big |^2}{|f_1^{(m)}|^4}\right) . \end{array}\right. } \end{aligned}$$
(3.17)

Now, we prove that if \(\varphi :S^2 \rightarrow HP^n\) is a linearly full quaternionic mixed pair with parallel second fundamental form, then, up to \(Sp(n+1),\) it belongs to one of the following three cases: (2) \(\varphi =\left[ (1,~z)^T\right] :S^2\rightarrow CP^1\subset HP^1\), with \(K=2, ~B=0;\)

(3) \(\varphi =\left[ (1,~\sqrt{2}z,z^2)^T\right] :S^2\rightarrow CP^2\subset HP^2\), with \(K=1, ~||B||^2=2\);

(4) \(\varphi =\left[ \!(1-\frac{1}{2}\overline{z}^3j,~\sqrt{3}z+\frac{\sqrt{3}}{2}\overline{z}^2j,\frac{3}{2}z^2,\frac{\sqrt{3}}{2}z^3)^T\!\right] :S^2\rightarrow \textit{HP}^3,\) with \(K=\frac{2}{3}, ~||B||^2=\frac{1}{3}\).

If \(\varphi \) is a linearly full quaternionic mixed pair with parallel second fundamental form, then applying Lemma 3.1 and substituting (3.17) into (3.1), we get

$$\begin{aligned} \frac{\Big |\left\langle f_2^{(m)}, ~\mathbf {J}{f_1^{(m)}}\right\rangle \Big |^2}{|f_1^{(m)}|^4}=\frac{3}{4}l_1 ^{(m)}\left( \frac{l_1 ^{(m)}}{l_0^{(m)}}-1\right) . \end{aligned}$$
(3.18)

Since the metric \(ds^2=2l_{0}^{(m)}dzd\overline{z}\) induced by \(\varphi \) is of constant curvature, and the metric induced by \(\underline{f}_0^{(m)}\) is \(ds^2=l_{0}^{(m)}dzd\overline{z}\), then it follows from ([2], Theorem 5.4) that \(\underline{f}_0^{(m)}, \underline{f}_1^{(m)}, \ldots , \underline{f}_{m}^{(m)}\) is the Veronese sequence in \(CP^{m}\subset CP^{2n+1}\), up to \(U(2n+2)\). Then from (3.7) and (3.18) we get

$$\begin{aligned} \Big |\left\langle f_2^{(m)}, ~\mathbf {J}{f_1^{(m)}}\right\rangle \Big |^2=\frac{3m(m-1)(m-2)}{2}(1+z\overline{z})^{2m-6}. \end{aligned}$$
(3.19)

We denote by \(r\) the isotropy order of \(\varphi \) (cf. [3], §3A). If \(r\) is finite, then \(r=2s~(1\le s\le n+1)\) by ([1], Proposition 3.2). Otherwise, \(r = \infty \), in which case \(\varphi \) is called strongly isotropic (cf. [1], section 2C).

If \(m=1\), observing (3.17), we find \(K=2, ~B=0\). It belongs to the case of totally geodesic. If \(m=2\), since \(r\ge 2\), which implies \({\underline{f}}_2^{(2)} \perp {\mathbf {J}}{\underline{f}_0^{(2)}}\) by (3.16), then we have \(\left\langle f_2^{(2)}, ~\mathbf {J}{f_1^{(2)}}\right\rangle =\partial \left\langle f_2^{(2)}, ~\mathbf {J}{f_0^{(2)}}\right\rangle =0\), which implies (3.19) holds. Hence, its second fundamental form is parallel. In fact, the above two cases are both strongly isotropic.

If \(m\ge 3\), from (3.19) we find \(\left\langle f_3^{(m)}, ~\mathbf {J}{f_0^{(m)}}\right\rangle =-\left\langle f_2^{(m)}, ~\mathbf {J}{f_1^{(m)}}\right\rangle \ne 0\), which implies in this case \(r=2\). In the following, we discuss the above three cases, respectively.

Case Ib1, \(m=1\).

In this case, we have

$$\begin{aligned} \underline{\varphi }_0=U\underline{V}_0^{(1)} \oplus {\mathbf {J}} U\underline{V}_0^{(1)}, \end{aligned}$$
(3.20)

where \(\underline{V}_0^{(1)}\) is a Veronese surface in \(CP^1\subset CP^{2n+1}\) with the standard expression given in ([2], section 5), and \(U\in U(2n+2)\) satisfies \(tr\Big (V_1^{(1)} {V_0^{(1)}}^{T} U^{T}J_{n+1}U\Big )=0\), as this expresses the orthogonality of \(Jf_0^{(1)}\) and \(f_1^{(1)}\).

Similarly, we get the type of \(\overline{W}=U^{T}J_{n+1}U\in U(2n+2)\) as follows:

(3.21)

As in case (1a), by (3.12),(3.13), and (3.21), we get

$$\begin{aligned} {\left\{ \begin{array}{ll} e_2=\overline{W}\overline{e}_1=\left( 0,0,-a_{13},-a_{14},\ldots ,-a_{1,2n+2}\right) ^T,\\ e_4=\overline{W}\overline{e}_3=\left( 0,0,-a_{23},-a_{24},\ldots ,-a_{2,2n+2}\right) ^T. \end{array}\right. } \end{aligned}$$

Since \(\overline{W}\) in (3.21) is a unitary matrix, \(\{e_1,e_2,e_3,e_4\}\) are mutually orthogonal. Similarly, we can choose \(n-1\) proper unit column vectors \(e_{2p+1}=(0,0,*)^T\in \mathbb {C}^{2n+2}~(2\le p\le n)\) such that \(\{e_1,e_2,\ldots ,e_{2n+1},e_{2n+2}=\overline{W}\overline{e}_{2n+1}\}\) are mutually orthogonal, and the type of the corresponding \(U\) is as follows:

(3.22)

Thus, we have

$$\begin{aligned}&UV_0^{(1)}=(1,~0,~z,~0,~0,\ldots ,0)^T,\\&{\mathbf {J}}UV_0^{(1)}=(0,1,~0,~\overline{z},~0,\ldots ,0)^T. \end{aligned}$$

Obviously, in this case \(\varphi \) is congruent to the case (2) with \(K=2,~B=0\).

Case Ib2, \(m=2\).

In this case, we have

$$\begin{aligned} \underline{\varphi }_0=U\underline{V}_0^{(2)} \oplus {\mathbf {J}} U\underline{V}_0^{(2)}, \end{aligned}$$
(3.23)

where \(\underline{V}_0^{(2)}\) is a Veronese surface in \(CP^2\subset CP^{2n+1}\) with the standard expression given in ([2], §5), and \(U\in U(2n+2)\) satisfies \(tr\Big ( V_1^{(2)} {V_0^{(2)}}^{T} U^{T}J_{n+1}U\Big )=0\), as this expresses the orthogonality of \(Jf_0^{(2)}\) and \(f_1^{(2)}\).

Similarly, we get the type of \(\overline{W}=U^{T}J_{n+1}U\in U(2n+2)\) as follows:

(3.24)

And the type of the corresponding \(U\) is as follows:

(3.25)

Thus, we have

$$\begin{aligned} UV_0^{(2)}=(1,~0,~\sqrt{2}z,~0,~z^2,~0,~0,\ldots ,0)^T,\\ {\mathbf {J}}UV_0^{(2)}=(0,~1,~0,~\sqrt{2}\overline{z},~0,~\overline{z}^2,~0,\ldots ,0)^T. \end{aligned}$$

Obviously, in this case \(\varphi \) is congruent to the case (3) with \(K=1,~\Vert B \Vert ^2=2\).

Case Ib3, \(m\ge 3\).

In this case, the trivial bundle \(S^2 \times \mathbb {C}^{2n+2}\) over \(S^2\) has a corresponding decomposition \(S^2 \times \mathbb {C}^{2n+2} = S^2 \times \mathbb {C}^{m+1} \oplus S^2 \times \mathbb {C}^{2n-m+1}\). From (3.16) we set \(\mathbf {J}{f_0^{(m)}}=x_3f_3^{(m)}+x_4f_4^{(m)}+\cdots +x_mf_m^{(m)}+V\), where \(x_i~(i=3,\ldots ,m)\) are complex coefficients and bundle \(\underline{V}\subset S^2 \times \mathbb {C}^{2n-m+1}\). Then, it follows from \(\partial {\mathbf {J}{f_0^{(m)}}}=0\) that

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial x_3+x_3\partial \log |f_3^{(m)}|^2=0,\\ \partial x_i+x_{i-1}+x_i\partial \log |f_i^{(m)}|^2=0,~(i=4,\ldots ,m),\\ \partial V=0. \end{array}\right. } \end{aligned}$$
(3.26)

And we have \(\left\langle f_3^{(m)}, ~\mathbf {J}{f_0^{(m)}}\right\rangle =\overline{x}_3 |f_3^{(m)}|^2\). Then, from ([2], §5) and (3.19) we get

$$\begin{aligned} |x_3|^2|f_3^{(m)}|^4=\frac{3m(m-1)(m-2)}{2}(1+z\overline{z})^{2m-6}. \end{aligned}$$
(3.27)

By (3.26) and (3.27), we find \(\partial \overline{\partial }\log \left( |x_3|^2|f_3^{(m)}|^4\right) =\frac{2m-6}{(1+z\overline{z})^2}=0\), which implies \(m=3\), i.e.,

$$\begin{aligned} \underline{\varphi }_0=U\underline{V}_0^{(3)} \oplus {\mathbf {J}} U\underline{V}_0^{(3)}, \end{aligned}$$
(3.28)

where \(\underline{V}_0^{(3)}\) is a Veronese surface in \(CP^3\subset CP^{2n+1}\) with the standard expression given in ([2], §5), and \(U\in U(2n+2)\) satisfies \(tr\Big ( V_1^{(3)} {V_0^{(3)}}^{T} U^{T}J_{n+1}U\Big )=0\) and \(tr\Big ( V_3^{(3)} {V_0^{(3)}}^{T} U^{T}J_{n+1}U\Big )\ne 0\), as these express \(r=2\).

Similarly, we get the type of \(\overline{W}=U^{T}J_{n+1}U\in U(2n+2)\) as follows:

(3.29)

where \(a_{14}\ne 0\).

In this case, if \(|a_{14}|^2=1\), as in the case (Ia) we choose (3.13), then choose \(n-1\) proper unit column vectors \(e_{2p+1}=(0,0,0,0,*)^T\in \mathbb {C}^{2n+2}~(2\le p\le n)\) such that the type of the corresponding \(U\in U(2n+2)\) is as follows:

(3.30)

Thus, we have

$$\begin{aligned}&UV_0^{(3)}=(1,~-a_{14}z^3,~\sqrt{3}z,~\sqrt{3}a_{14}z^2,~0,~0,\ldots ,0)^T,\\&{\mathbf {J}}UV_0^{(3)}=(\overline{a}_{14}\overline{z}^3,~1,~-\sqrt{3}\overline{a}_{14}\overline{z}^2,~\sqrt{3}\overline{z},~0,~0,\ldots ,0)^T. \end{aligned}$$

If \(|a_{14}|^2\ne 1\), then we choose

$$\begin{aligned} {\left\{ \begin{array}{ll} e_1=(1,0,0,0,0,\ldots ,0)^T,\\ e_3=(0,1,0,0,0,\ldots ,0)^T,\\ e_5=\frac{1}{\sqrt{1-|a_{14}|^2}}(0,0,1-|a_{14}|^2,0,\overline{a}_{14}a_{25},\ldots ,\overline{a}_{14}a_{2,2n+2})^T,\\ e_7=\frac{1}{\sqrt{1-|a_{14}|^2}}(0,0,0,1-|a_{14}|^2,-\overline{a}_{14}a_{15},\ldots ,-\overline{a}_{14}a_{1,2n+2})^T. \end{array}\right. } \end{aligned}$$
(3.31)

And we choose \(n-3\) proper unit column vectors \(e_{2p+1}=(0,0,0,0,*)^T\in \mathbb {C}^{2n+2}~(4\le p\le n)\) such that the type of the corresponding \(U\in U(2n+2)\) is as follows:

(3.32)

Thus, we have

$$\begin{aligned} UV_0^{(3)}{=}(1,~-a_{14}z^3,~\sqrt{3}z,~\sqrt{3}a_{14}z^2,\sqrt{3-3|a_{14}|^2}z^2,~0,\sqrt{1-|a_{14}|^2}z^3,~0,~0,\ldots ,0)^T,\\ {\mathbf {J}}UV_0^{(3)}{=}(\overline{a}_{14}\overline{z}^3,1,~-\sqrt{3}\overline{a}_{14}\overline{z}^2,~\sqrt{3}\overline{z},~0,\sqrt{3-3|a_{14}|^2}\overline{z}^2,~0,~\sqrt{1-|a_{14}|^2}\overline{z}^3,~0,\ldots ,0)^T. \end{aligned}$$

From (3.30) and (3.32), we have \(\left\langle f_3^{(3)}, ~\mathbf {J}{f_0^{(3)}}\right\rangle =\left\langle UV_3^{(3)}, ~\mathbf {J}{U V_0^{(3)}}\right\rangle =6a_{14}\). On the other hand, from (3.27) we get \(\left| \left\langle f_3^{(3)}, ~\mathbf {J}{f_0^{(3)}}\right\rangle \right| ^2=9\). So \(a_{14}=\frac{1}{2}e^{\sqrt{-1}\theta }~(0\le \theta \le 2\pi )\). Hence, in this case \(\varphi \) is congruent to the case (4) with \(K=\frac{2}{3},~\Vert B \Vert ^2=\frac{1}{3}\).

(II) Let \(\varphi :S^2 \rightarrow HP^n\) be an irreducible linearly full harmonic map. At first, we state a conclusion about parallel minimal immersions of 2-spheres in \(G(k,N)\) as follows:

Lemma 3.2

([7]) Let \(\varphi :S^2 \rightarrow G(k,N)\) be a conformal minimal immersion with the second fundamental form \(B\). Suppose that \(B\) is parallel, then the following equations

$$\begin{aligned} \left\{ \begin{array}{l} \lambda ^2(2K+||B||^2)A_{z}^*+4\left[ [A_z,A_z^*],A_{z}^*\right] =0,\\ \lambda ^2(\frac{||B||^2}{4}-K)P+\left[ [A_z,A_z^*],P\right] =0 \end{array}\right. \end{aligned}$$
(3.33)

hold.

By [1], we know \(\varphi _0\) belongs to the following harmonic sequence

$$\begin{aligned} 0{\longleftarrow } \cdots \mathop {\longleftarrow }\limits ^{A''_{\varphi _2}} \underline{\varphi }_{-2} \mathop {\longleftarrow }\limits ^{A''_{\varphi _1}} \underline{\varphi }_{-1} \mathop {\longleftarrow }\limits ^{A''_{\varphi _0}} \underline{\varphi }_0 \mathop {\longrightarrow }\limits ^{A'_{\varphi _0}} \underline{\varphi }_{1} \mathop {\longrightarrow }\limits ^{A'_{\varphi _1}} \underline{\varphi }_{2} \mathop {\longrightarrow }\limits ^{A'_{\varphi _2}} \cdots {\longrightarrow } 0, \end{aligned}$$
(3.34)

where \(\underline{\varphi }_0=\underline{i_2\circ \varphi },~\underline{\varphi }_{-1}={\mathbf {J}}\underline{\varphi }_{1},~\underline{\varphi }_{-2}={\mathbf {J}}\underline{\varphi }_{2}\).

We choose a unit column vector \(X\in \underline{\varphi }_0\), then we have

$$\begin{aligned} \underline{\varphi }_0=\underline{X} \oplus \mathbf {J} \underline{X},\quad \underline{\varphi }_1=span\Big \{\underline{X}_1, \underline{Y}_1\Big \}, \end{aligned}$$
(3.35)

where \(X_1=\partial {X}-\left\langle \partial {X},X\right\rangle X-\left\langle \partial {X},\mathbf {J}X\right\rangle \mathbf {J}X\) and \(Y_1=\partial {\mathbf {J}X}-\left\langle \partial {\mathbf {J}X},X\right\rangle X-\left\langle \partial {\mathbf {J}X},\mathbf {J}X\right\rangle \mathbf {J}X\). Here, \(X_1\) and \(Y_1\) are not orthogonal in general.

Let

$$\begin{aligned} E=[X,\mathbf {J}X]^*\partial [X,\mathbf {J}X]. \end{aligned}$$
(3.36)

Then, from (3.35), we have

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial [X,\mathbf {J}X]=[X,\mathbf {J}X]E+[X_1,Y_1],\\ \overline{\partial }[X,\mathbf {J}X]=-[X,\mathbf {J}X]E^*+[-\mathbf {J}Y_1,\mathbf {J}X_1], \end{array}\right. } \end{aligned}$$
(3.37)

where \([X_1,Y_1]^*[X,\mathbf {J}X]=[-\mathbf {J}Y_1,\mathbf {J}X_1]^*[X,\mathbf {J}X]=0\).

From (3.37) and the identity \(\partial \overline{\partial }=\overline{\partial }\partial \), we get

(3.38)

From (3.35), we have \(\varphi _0=XX^*+({\mathbf {J}} X)({\mathbf {J}} X)^*\). Then by (2.1), a straightforward calculation shows

$$\begin{aligned} {\left\{ \begin{array}{ll} \lambda ^2=2(|X_1|^2+|Y_1|^2),\\ A_{z}=(\mathbf {J}X)(\mathbf {J}X_1)^*-X(\mathbf {J}Y_1)^*-X_1X^*-Y_1(\mathbf {J}X)^*. \end{array}\right. } \end{aligned}$$
(3.39)

Since \(\varphi _0\) is harmonic, the corresponding equivalent condition \(\overline{\partial }A_z+A_zA_z^*-A_z^*A_z=0\) (cf. [12]) implies

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\left\langle \overline{\partial }X_1,X_1\right\rangle }{|X_1|^2}=-\frac{\left\langle \overline{\partial }Y_1,Y_1\right\rangle }{|Y_1|^2}=\left\langle \overline{\partial }X,X\right\rangle ,\\ \frac{\left\langle \overline{\partial }X_1,Y_1\right\rangle }{|Y_1|^2}=\left\langle \overline{\partial }X,\mathbf {J}X\right\rangle ,\quad \frac{\left\langle \overline{\partial }Y_1,X_1\right\rangle }{|X_1|^2}=\left\langle \overline{\partial }\mathbf {J}X,X\right\rangle . \end{array}\right. } \end{aligned}$$
(3.40)

Now, we prove that if \(\varphi :S^2 \rightarrow HP^n\) is an irreducible linearly full harmonic map with parallel second fundamental form, then, up to \(Sp(n+1),\) it belongs to one of the following two cases: (5) \(\varphi =\left[ (-2\overline{z},~\sqrt{2}-\sqrt{2}z\overline{z},~2z)^T\right] :S^2\rightarrow CP^2\subset HP^2\), with \(K=\frac{1}{2}, ~B=0;\) (6) \(\varphi =\left[ (6\overline{z}^2,~-6\overline{z}+6z\overline{z}^2,~\sqrt{6}-4\sqrt{6}z\overline{z}+\sqrt{6}z^2\overline{z}^2,~6z-6z^2\overline{z}, ~6z^2)^T\right] :S^2\rightarrow CP^4\subset HP^4,\) with \(K=\frac{1}{6}, ~\Vert B\Vert ^2=\frac{2}{3}\).

If \(\varphi :S^2 \rightarrow HP^n\) is an irreducible linearly full harmonic map with parallel second fundamental form, then applying Lemma 3.2 and substituting (3.39) into the first equation of (3.33), we get

$$\begin{aligned} \lambda ^2=4|X_1|^2,\quad 2K+||B||^2=1,\quad \left\langle X_1,Y_1\right\rangle =0,\quad |X_1|^2=|Y_1|^2. \end{aligned}$$
(3.41)

From (3.38) and (3.41) we have

$$\begin{aligned} \overline{\partial }E+\partial E^*+[E,E^*]=0. \end{aligned}$$
(3.42)

Let \(\widetilde{X}\in \underline{\varphi }_0\) be another unit column vector such that \(\underline{\varphi }_0=\underline{\widetilde{X}} \oplus \mathbf {J} \underline{\widetilde{X}}\), then

$$\begin{aligned}{}[\widetilde{X},\mathbf {J}\widetilde{X}]=[X,\mathbf {J}X]T, \end{aligned}$$
(3.43)

where \(T:S^2\rightarrow SU(2)\) is to be determined such that \(\widetilde{X}\) satisfies \([\widetilde{X},\mathbf {J}\widetilde{X}]^*d[\widetilde{X},\mathbf {J}\widetilde{X}]=0\). Such \(T\) is a solution of the linear PDE

$$\begin{aligned} dT+(Edz-E^*d\overline{z})T=0. \end{aligned}$$
(3.44)

The integrability condition of (3.44) is just (3.42), so it has a unique solution locally on \(S^2\) for any given initial value. Let \(T\) be a solution of (3.44) with the initial value \(T(0)\in SU(2)\). From (3.44) we have \(d(T^*T)=0\) and \(d|T|=0\), so \(T\in SU(2)\).

Now, we choose a unit column vector \(X\in \underline{\varphi }_0\) such that \(\underline{\varphi }_0=\underline{X} \oplus \mathbf {J} \underline{X}\) and

$$\begin{aligned}{}[X,\mathbf {J}X]^*d[X,\mathbf {J}X]=0. \end{aligned}$$
(3.45)

It follows from (3.40) and (3.45) that

$$\begin{aligned} \partial \overline{\partial }X=-|X_1|^2X. \end{aligned}$$
(3.46)

Let \(\underline{f}=[X]:S^2\rightarrow CP^{2n+1}\) be a smooth immersion. Similarly, by calculating the equivalent condition of harmonic, we find \(\underline{f}\) is harmonic by (3.45) and (3.46). Of course, \(\mathbf {J}\underline{f}=[\mathbf {J}X]:S^2\rightarrow CP^{2n+1}\) is also harmonic. So \(\underline{\varphi }_0=\underline{f}\oplus \mathbf {J}\underline{f}\), where \(\underline{f}\) belongs to the following harmonic sequence

$$\begin{aligned} 0{\longrightarrow } \cdots \mathop {\longrightarrow }\limits ^{A_{p-1}''} \underline{f}_{p-1} \mathop {\longrightarrow }\limits ^{A_{p}''} \underline{f}_p=\underline{f} \mathop {\longrightarrow }\limits ^{A_{p}'} \underline{f}_{p+1} \mathop {\longrightarrow }\limits ^{A_{p+1}'} \cdots {\longrightarrow } 0. \end{aligned}$$
(3.47)

As in the case (Ia), let \(f_0\) be a holomorphic section of \({\underline{f}} _0\), i.e. \(\overline{\partial } f_0 = 0\), and \(f_p\) satisfy the corresponding formulas. From (3.41), we know \(l_{p-1}=l_p\), which implies that \(\underline{f}_p\) is totally real by ([2], Theorem 7.3), i.e. \(\underline{f}_p=\underline{f}_m^{(2m)}:S^2\rightarrow RP^{2m}\subset CP^{2m}\subset CP^{2n+1}\), where \(2\le 2m\le 2n+1\). Let \(f_p={f}_m^{(2m)}\) satisfy the corresponding formulas, then in harmonic sequence (3.34) by (3.40) and (3.45) we have

$$\begin{aligned} \underline{\varphi }_0=\underline{f}_m^{(2m)} \oplus {\mathbf {J}} \underline{f}_m^{(2m)},\quad \underline{\varphi }_1=\underline{f}_{m+1}^{(2m)} \oplus {\mathbf {J}} \underline{f}_{m-1}^{(2m)},\quad \underline{\varphi }_2=\underline{f}_{m+2}^{(2m)}\oplus {\mathbf {J}} \underline{f}_{m-2}^{(2m)}, \end{aligned}$$
(3.48)

where \(l_i^{(2m)}=l_{2m-1-i}^{(2m)}\ (i=0,\ldots ,m-1)\) and \(\underline{\varphi }_0,\ \underline{\varphi }_1,\ \underline{\varphi }_2\) are mutually orthogonal.

At this time, from (3.48), we have \(\varphi _0=\frac{{f}_m^{(2m)}({f}_m^{(2m)})^*}{|{f}_m^{(2m)}|^2}+\frac{({\mathbf {J}} {f}_m^{(2m)})({\mathbf {J}} {f}_m^{(2m)})^*}{|{f}_{m}^{(2m)}|^2}\). Then by (2.1) and a series of calculations, we obtain

$$\begin{aligned} \left\{ \begin{array}{l} \lambda ^2=4l_m^{(2m)},\\ K=\frac{1}{2}-\frac{l_{m+1}^{(2m)}}{2l_{m}^{(2m)}},\\ A_{z}=\frac{(\mathbf {J}f_{m}^{(2m)})(\mathbf {J}f_{m+1}^{(2m)})^*}{|f_{m}^{(2m)}|^2}+\frac{(\mathbf {J}f_{m-1}^{(2m)})(\mathbf {J}f_{m}^{(2m)})^*}{|f_{m-1}^{(2m)}|^2} -\frac{f_{m+1}^{(2m)}{f_{m}^{(2m)}}^*}{|f_{m}^{(2m)}|^2}-\frac{f_{m}^{(2m)}{f_{m-1}^{(2m)}}^*}{|f_{m-1}^{(2m)}|^2},\\ P=\frac{1}{4}\left[ \frac{l_{m-2}^{(2m)}}{|f_{m}^{(2m)}|^2}f_{m}^{(2m)}{f_{m-2}^{(2m)}}^*+\frac{(\mathbf {J}f_{m}^{(2m)})(\mathbf {J}f_{m+2}^{(2m)})^*}{|f_{m+1}^{(2m)}|^2} -\frac{f_{m+2}^{(2m)}{f_{m}^{(2m)}}^*}{|f_{m+1}^{(2m)}|^2}-\frac{l_{m-2}^{(2m)}}{|f_{m}^{(2m)}|^2}(\mathbf {J}f_{m-2}^{(2m)})(\mathbf {J}f_{m}^{(2m)})^*\right] ,\\ ||B||^2=\frac{l_{m+1}^{(2m)}}{l_{m}^{(2m)}}. \end{array}\right. \quad \end{aligned}$$
(3.49)

Then applying Lemma 3.2 and substituting (3.49) into the second equation of (3.33), we get \(m=1\) or

$$\begin{aligned} \left\langle \mathbf {J}f_{m-2}^{(2m)},f_{m-1}^{(2m)}\right\rangle =\left\langle f_{m+2}^{(2m)},\mathbf {J}f_{m+1}^{(2m)}\right\rangle =0,\ 3l_{m+1}^{(2m)}=2l_{m}^{(2m)}. \end{aligned}$$
(3.50)

In the latter case, since \(l_{m+1}^{(2m)}=\frac{(m+2)(m-1)}{(1+z\overline{z})^2}\) and \(l_{m}^{(2m)}=\frac{(m+1)m}{(1+z\overline{z})^2}\) by ([2], §5), we have \(m=2\) by (3.50). Hence, in the following, we discuss the above two cases of \(m=1\) and \(m=2\) respectively.

Case II1, \(m=1\).

In this case, by (3.50) we have

$$\begin{aligned} \underline{\varphi }_0=U\underline{V}_1^{(2)} \oplus {\mathbf {J}} U\underline{V}_1^{(2)}, \end{aligned}$$
(3.51)

where \(\underline{V}_1^{(2)}\) is a Veronese surface in \(CP^2\subset CP^{2n+1}\) with the standard expression given in ([2], Section 5) and \(U\in U(2n+2)\) satisfies \(tr\Big ( V_2^{(2)} {V_0^{(2)}}^{T} U^{T}J_{n+1}U\Big )=0\), as this expresses the orthogonality of \(Jf_0^{(2)}\) and \(f_2^{(2)}\).

By calculating, we find in this case \(\overline{W}=U^{T}J_{n+1}U\in U(2n+2)\) is the same type as (3.24). Then, the type of the corresponding \(U\in U(2n+2)\) is the same as (3.25). Thus, we have

$$\begin{aligned}&UV_1^{(2)}=\frac{1}{1+z\overline{z}}(-2\overline{z},~0,~\sqrt{2}-\sqrt{2}z\overline{z},~0,~2z,~0,~0,\ldots ,0)^T,\\&{\mathbf {J}}UV_1^{(2)}=\frac{1}{1+z\overline{z}}(0,~-2{z},~0,~\sqrt{2}-\sqrt{2}z\overline{z},~0,~2\overline{z},~0,\ldots ,0)^T. \end{aligned}$$

In this case, it is easy to check that the corresponding map \(\varphi \) is totally geodesic. Obviously, it is congruent to the case (5) with \(K=\frac{1}{2},~B=0\).

Case II2, \(m=2\).

In this case, by (3.50) we have

$$\begin{aligned} \underline{\varphi }_0=U\underline{V}_2^{(4)} \oplus {\mathbf {J}} U\underline{V}_2^{(4)}, \end{aligned}$$
(3.52)

where \(\underline{V}_2^{(4)}\) is a Veronese surface in \(CP^4\subset CP^{2n+1}\) with the standard expression given in ([2], Section 5) and \(U\in U(2n+2)\) satisfies \(tr\Big ( V_4^{(4)} {V_0^{(4)}}^{T} U^{T}J_{n+1}U\Big )=0\), as this expresses the orthogonality of \(Jf_0^{(4)}\) and \(f_4^{(4)}\).

Similarly, we get the type of \(\overline{W}=U^{T}J_{n+1}U\in U(2n+2)\) as follows:

And the type of the corresponding \(U\in U(2n+2)\) is as follows:

(3.53)

Thus, we have

$$\begin{aligned} UV_2^{(4)}&= \frac{2}{(1+z\overline{z})^2}(6\overline{z}^2,0,-6\overline{z}+6z\overline{z}^2,0,\sqrt{6}-4\sqrt{6}z\overline{z} \\&+\,\,\sqrt{6}z^2\overline{z}^2,0,6z-6z^2\overline{z},0,6z^2,0,\ldots ,0)^T,\\ {\mathbf {J}}UV_2^{(4)}&= \frac{2}{(1+z\overline{z})^2}(0,~6{z}^2,0,-6z+6z^2\overline{z},0,\sqrt{6}-4\sqrt{6}z\overline{z} \\&+ \,\,\sqrt{6}z^2\overline{z}^2,0,6\overline{z}-6z\overline{z}^2,0,6\overline{z}^2,\ldots ,0)^T. \end{aligned}$$

In this case, it is easy to check that the Eq. (3.1) holds, which shows the second fundamental form of the corresponding map \(\varphi \) is parallel. Obviously, it is congruent to the case (6) with \(K=\frac{1}{6},~||B||^2=\frac{2}{3}\).

It is easy to check that no two of the above six cases are congruent, i.e., we cannot transform any one into any other by left multiplication by \(Sp(n+1)\). To sum up, we get Theorem 1.1.