You are on page 1of 8

mathematics

Article
The Harmonic Mapping Whose Hopf Differential
Is a Constant
Liang Shen
School of Mathematics and Statistics, Beijing Institute of Technology, Beijing 100081, China; shenl@bit.edu.cn

Received: 16 July 2020; Accepted: 3 August 2020; Published: 7 August 2020 

Abstract: Suppose that h(z) is a harmonic mapping from the unit disk D to itself with respect to the
hyperbolic metric. If the Hopf differential of h(z) is a constant c > 0, the Beltrami coefficient µ(z) of
h(z) is radially symmetric and takes the maximum at z = 0. Furthermore, the mapping γ : c → µ(0)
is increasing and gives a homeomorphism from (0, +∞) to (0, 1).

Keywords: harmonic mapping; Hopf differential; Beltrami coefficient

1. Introduction
Let h(z) be an orientation preserving diffeomorphism from the unit disk D to itself. The mapping
h(z) is harmonic with respect to the hyperbolic metric on D if it satisfies the Euler–Lagrange equation

ρz ( h)
hzz̄ + 2 hz hz̄ = 0, (1)
ρ(h)

where ρ2 (z)|dz|2 = (1−|4z|2 )2 |dz|2 is the hyperbolic metric on D.


The harmonic mapping is closed related to the quasiconformal mapping and Teichmüller theory.
For example, the Schoen conjecture, proved by Li and Tam in [1,2] and Markovic in [3], says that for
every quasisymmetric homeomorphism on the unit circle T, there exists a unique harmonic extension
to the interior D.
The Hopf differential φ(z) = ρ2 (h)hz hz̄ of a harmonic mapping h(z) is always holomorphic by (1).
Given a Hopf differential, the existence and uniqueness of the harmonic mapping is established by
Wan in [4]. Also in [4], it is proved that a harmonic mapping is quasiconformal if and only if the Bers
norm of the Hopf differential is bounded. Especially if the Hopf differential is constant, the harmonic
mapping is quasiconformal. In the paper, we consider such mappings and prove the following result.

Theorem 1. Suppose h(z) is a harmonic mapping whose Hopf differential is a constant c > 0. The Beltrami
coefficient µ(z) of h(z) is a radially symmetric real function and takes the maximum at z = 0. The mapping
γ : c → µ(0) is increasing and gives a homeomorphism from (0, +∞) to (0, 1).

Let hc (z) be a harmonic mapping with the Hopf differential cφ(z) between compact surfaces.
In [5], Wolf proved that the Beltrami coefficient µc (z) of hc (z) is an increasing function of c > 0,
which is used to give a compactification of Teichmüller space. But the method can not be generalized
to the harmonic mapping on the unit disk. In the proof of Theorem 1, we use the similar method as
in [4,6], i.e., the generalized maximum principle of Cheng and Yau [7].
In the following, the Laplace–Beltrami operator is denoted by ∆ g on a Riemannian manifold
( M, g). For a conformal metric σ|dz|2 on the unit disk D, the Laplace–Beltrami operator has the
following form

∆σ = ,
σ

Mathematics 2020, 8, 1310; doi:10.3390/math8081310 www.mdpi.com/journal/mathematics


Mathematics 2020, 8, 1310 2 of 8

∂2 ∂2
where ∆ = ∂x2
+ ∂y2
is the Laplace operator.

2. The Absolute Value of the Beltrami Coefficient Is Radially Symmetric


First we give some known facts about harmonic mappings.
Let h(z) be a harmonic mapping with respect to the hyperbolic metric on the unit disk D. Let φ(z)
be the Hopf differential and µ(z) be the Beltrami coefficient of h(z). The Bochner formula gives:

¯ |2 − 1,
∆ρ2 log |∂h| = |∂h|2 − |∂h ¯ | = −(|∂h|2 − |∂h
∆ρ2 log |∂h ¯ |2 ) − 1

ρ(h(z)) ¯ ρ(h(z))
where ∂h = ρ(z)
hz , ∂h = ρ(z)
hz̄ . By the first equality, we have

∆v = e2v − |φ|2 e−2v and ∆v > 0.

where e2v = ρ2 (h(z))|hz |2 . The Beltrami coefficient µ(z) satisfies that

hz̄ ρ2 (h(z))hz hz̄ φ̄


µ(z) = = 2 = 2v .
hz ρ (h(z))|hz |2 e

On the other hand, the following theorem was proved by Wan in [4].

Theorem 2 (Theorem 12, [4]). For any holomorphic function φ(z) on D, there exists a unique smooth function
v(z) such that

1. ∆v = e2v − |φ|2 e−2v > 0,


2. e2v |dz|2 is complete.

Theorem 2 gives a way to construct a harmonic mapping from a Hopf differential. From the
Hopf differential φ one has the Beltrami coefficient µ(z), then solve the Beltrami equation to obtain the
harmonic mapping.
In the following, we discuss the harmonic mappings which has special Hopf differentials. First we
show for a class of Hopf differentials, the metric e2v |dz|2 is radially symmetric.

Lemma 1. Suppose the Hopf differential φ(z) is czn where c ∈ C and n = 0, 1, 2, · · · . The metric e2v |dz|2 is
radially symmetric, i.e., v(eiθ z) = v(z).

Proof. For the Hopf differential φ = czn , by Theorem 2, there is a unique smooth function v(z)
such that

1. ∆v = e2v − |φ|2 e−2v > 0,


2. e2v |dz|2 is complete.

Let w = eiθ z and v1 (z) = v(eiθ z). Since

∆v(w) = e2v(w) − |φ(w)|2 e−2v(w)

and
∂2 v2 ∂2
∆v(w) = 4 = 4 v2 (eiθ z) = ∆(v(eiθ z)),
∂w ∂w̄ w=eiθ z ∂z ∂z̄
we have
∆v1 = e2v1 − |czn |2 e−2v1 .

By the uniqueness of the solution, the function v1 (z) must be v(z). So v(z) is radially symmetric.
Mathematics 2020, 8, 1310 3 of 8

In particular, if the Hopf differential is a constant c ∈ C, the metric e2v |dz|2 is


radially symmetric. If φ1 (z) and φ2 (z) are two holomorphic functions which satisfy that φ1 (z) =
eiα φ2 (z), the corresponding v1 (z) and v2 (z) satisfy v1 (z) = v2 (z) by the uniqueness. As a result,
the corresponding Beltrami coefficients satisfies that µ1 (z) = e−iα µ2 (z). So we only need consider the
Hopf differential is a positive number c > 0. In these cases, the Beltrami coefficients are real functions.
In the following lemma, we consider the maximum of |µ(z)|.

Lemma 2. Let φ(z) be Hopf differential of a harmonic mapping h(z) and µ(z) be the Beltrami coefficient.
If φ(z) has no zeros, the function |µ(z)| has no minimum in D.

φ
Proof. Since |φ(z)| > 0 and µ = e2v
, the Beltrami coefficient µ(z) satisfies that |µ(z)| > 0.
The Bochner formula

¯ |2 − 1,
∆ρ2 log |∂h| = |∂h|2 − |∂h ¯ | = −(|∂h|2 − |∂h
∆ρ2 log |∂h ¯ |2 ) − 1

hold for every z ∈ D. Then we have

∆ρ2 log |µ(z)| = −2|∂h|2 (1 − |µ(z)|2 ), (2)

which implies that ∆ log |µ(z)| < 0 holds for every z ∈ ∆. So log |µ(z)| is a superharmonic function
on D, which implies that log |µ(z)| can not obtain the minimum in D. Then the function |µ(z)| has no
minimum in D.

If the Hopf differential of a harmonic mapping h(z) is a constant c ∈ C, the absolute value of the
Beltrami coefficient µ is radially symmetric by Lemma 1. Since lim |µ(z)| = 0, the function |µ(z)| has
|z|→1
an absolute maximum in D. If |µ(z)| obtains a maximum at z0 6= 0, |µ(z)| also obtains a maximum
at eiθ z0 . As a result, |µ(z)| have a minimum in |z| < |z0 |, which contradicts to Lemma 2. So |µ(z)|
takes the absolute maximum only at z = 0. It is the context of the following lemma. Here we give a
direct proof.

Lemma 3. Suppose the Hopf differential of a harmonic mapping h(z) is a constant c ∈ C. The maximum of
|µ(z)| is obtained at z = 0.

Proof. From (2), the following equality holds

∆ log |µ(z)| = −2ρ2 (h(z))|hz |2 (1 − |µ(z)|2 ),

which yields
∆ log |µ(z)| = −2e2v (1 − |µ(z)|2 ).

Then we have
1 |φ|
− ∆ log |µ(z)| = (1 − |µ(z)|2 ).
2 |µ|
1
Define a function τ (z) = 2 log |µ1 | > 0. The above equality becomes

∆τ = (e2τ − e−2τ )|φ|.

Recall that the Laplace operator in polar coordinates is given by

∂2 1 ∂ 1 ∂2
∆= 2
+ + 2 2.
∂r r ∂r r ∂θ
Mathematics 2020, 8, 1310 4 of 8

Since τ (z) is radially symmetric, it is a function of r = |z| and satisfies

1
τ 00 (r ) + τ 0 (r ) = |c|(e2τ − e−2τ ) > 0
r
(rτ 0 (r ))0
The left side is equal to r , so (rτ 0 (r ))0 > 0, which implies rτ 0 (r ) is a strictly increasing
function of r > 0. The function τ ( x ) on the line (−1, 1) is an even function of x. It satisfies that
τ 0 (0) = 0.
So rτ 0 (r ) takes value 0 at z = 0. As a result, rτ 0 (r ) > 0, which means τ (r ) is a strictly
increasing function. Then |µ(z)| takes maximum at z = 0.

Finally we give a geometric description of a harmonic mapping whose Hopf differential is


a constant.

Lemma 4. Suppose the Hopf differential of a harmonic mapping h(z) is a constant c > 0. The following
properties hold:
(1). Suppose the mapping h(z) satisfies that h(0) = 0, h(1) = 1, then h(z) maps the x-axis onto x-axis
and the y-axis onto y-axis.
(2). Let γ1 and γ2 be a pair of intersecting lines in the disk D which are horizontal and vertical respectively.
Then the curves h(γ1 ) and h(γ2 ) are orthogonal.

Proof. (1). The mapping H (z) = h(z̄) is also harmonic and has the same Hopf differential with h(z).
Since H (z) keeps 0, 1 fixed, we have H (z) = h(z). So h(z) maps the x-axis onto x-axis. Similarly we
have h(z) = −h(−z), which implies that h(z) maps the y-axis onto y-axis.
Now we prove (2). First assume that the Hopf differential of a harmonic mapping h(z) =
u( x, y) + iv( x, y) is a real number. By the following calculations,

1 1
hz = (h x − ihy ) = (u x + vy − i (uy − v x )),
2 2
1 1
hz̄ = (h x + ihy ) = (u x − vy + i (uy + v x )),
2 2
1
Im(φ) = ρ2 (h)Im(hz hz̄ ) = − ρ2 (h)(u x uy + v x vy ),
2
we find u x uy + v x vy = 0, which implies that the curves h(γ1 ) and h(γ2 ) are orthogonal.

3. The Value µc (0) Is Strictly Increasing


Let’s recall the following lemma of Cheng and Yau [8].

Lemma 5. Let ( M, g) be a complete manifold with the Ricci curvature bounded below by some constant.
Suppose α ∈ C2 ( M) satisfies
∆ g α ≥ f ( α ),

where f ( x ) is a continuous function which is positive and non-decreasing near ∞ and satisfies that
Z ∞ Z t − 12
dt f ( x )dx <∞
p q

for some constants p > q. Then α is bounded from above.

The following maximum principle given by Omori [9] and Yau [7] is well-known.
Mathematics 2020, 8, 1310 5 of 8

Lemma 6 (Omori–Yau maximum principle). Let ( M, g) be a complete manifold with the Ricci curvature of
M bounded below by some constant. If α ∈ C2 ( M) is bounded from above, then for any ε > 0, there is a point
xk ∈ M such that
α( xk ) > sup α − ε, |∇ g α( xk )| < ε, ∆ g α( xk ) < ε.
D

Let φ be a Hopf differential and e2v |dz|2 be the corresponding complete metric in Theorem 2.
The Gauss(Ricci) curvature of e2v |dz|2 is

∆v
K=− = |µ(z)|2 − 1.
e2v
Now we can prove the following lemma.

Lemma 7. Let φ1 and φ2 be Hopf differentials of harmonic mappings h1 , h2 on D respectively, which satisfy
that |φ1 (z)| = c > 0 and |φ1 | ≤ |φ2 |. Then the Beltrami coefficients µ1 and µ2 satisfy that |µ1 | ≤ |µ2 |.

Proof. Let τi (z) = 12 log |µ 1(z)| (i = 1, 2) as before. We have τi (z) > 0 since 0 < |µi (z)| < 1.
i
Define η (z) = τ2 (z) − τ1 (z). Since

∆τi = (e2τi − e−2τi )|φi | (i = 1, 2),

we have

∆η = |φ2 |(e2τ2 − e−2τ2 ) − |φ1 |(e2τ1 − e−2τ1 )


≥ c(e2τ2 − e−2τ2 ) − c(e2τ1 − e−2τ1 ).

By multiplying by e−2τ1 , the inequality becomes

e−2τ1 ∆η ≥ c(e2η − e−4τ1 e−2η − 1 + e−4τ1 ). (3)

Since e−4τ1 < 1, we have


e−2τ1 ∆η ≥ c(e2η − e−2η − 1). (4)

Let f ( x ) = c(e2x − e−2x − 1), which is positive and non-decreasing as x tends to +∞. The function
f ( x ) satisfies that
Z ∞ Z t − 12
dt f ( x )dx <∞
p q

for some constants p > q. The inequality (4) can be written as

∆ ρ1 η ≥ f ( η ).

1 e2v1
The metric ρ1 = e2τ1 is complete on D since e2τ1 = | µ1 |
= c .
The Gauss curvatures of e2τ1 and e2v1 satisfy

K (e2τ1 ) = cK (e2v1 ).

Since K (e2v1 ) = |µ1 |2 − 1 has a lower bound, the Gauss curvature of e2τ1 also has a lower bound.
Then we can use Lemma 5 to conclude that η is bounded from above. By Lemma 6, there is a sequence
of points xk in D, such that

lim η ( xk ) = sup η (= η̄ ), lim ∆ρ1 η ( xk ) ≤ 0.


k→∞ D k→∞
Mathematics 2020, 8, 1310 6 of 8

Since the sequence {e−4τ1 ( xk ) } is bounded, we can take a subsequence such that lim e−4τ1 ( xk ) =
k→∞
ξ (≤ 1). By taking limits on two sides of (3), we get

0 ≥ c(e2η̄ − ξe−2η̄ − 1 + ξ ).

Let
g(t) = c(e2t − ξe−2t − 1 + ξ ).

It is an increasing function of t and g(0) = 0. Then g(η̄ ) ≤ 0 implies that η̄ ≤ 0. So we have


η (z) ≤ 0 and τ2 ≤ τ1 .

To prove that γ is strictly increasing, we use the following lemma of Heinz [10].

Lemma 8. Let α(z) be a non-negative C2 -function on D( R) which satisfies

∆α ≤ Cα, C > 0.

Then for 0 ≤ r < R,


1 Cr2
ZZ
α(z)dσ ≤ e 4 α (0).
πr2 D(r )

Lemma 8 has the following corollary. It can also be proved directly. We sketch a proof for the sake
of completeness.

Corollary 1. Suppose α(z) ≥ 0 is defined on D( R) and α(0) = 0. If

∆α ≤ Cα, C > 0,

then α(z) ≡ 0.

Proof. Define Z 2π Z r
1
v (r ) = α(reiθ )dθ, V (r ) = v(ρ)dρ.
2π 0 0
By Green’s theorem,
Z 2π
1 d 1
ZZ
v 0 (r ) = α(reiθ )dθ = ∆αdσ,
2π dr 0 2πr D(r )

so Z r
1
ZZ
00 0 ρ
V (r ) − CV (r ) ≤ v (r ) − C v(ρ)dρ = (∆α − Cα)dσ ≤ 0.
0 r 2πr D(r )
√ √
Thus (V 02 (r))0 ≤ C(V 2 (r))0 , so V 0 (r) ≤ CV (r), i.e., (V (r)e Cr )0 ≤ 0, which implies V (r) ≡ 0.

Lemma 9. Let φ1 and φ2 be Hopf differentials of harmonic mappings h1 , h2 on D respectively, which satisfy
that |φ1 (z)| = c > 0 and |φ1 | < |φ2 |. Then |µ1 (z)| < |µ2 (z)| for all z ∈ D.

Proof. Recall that


∆τi = (e2τi − e−2τi )|φi |, (i = 1, 2)
Mathematics 2020, 8, 1310 7 of 8

where τi (z) = 12 log |µ 1(z)| (i = 1, 2). By Lemma 7, we have |µ1 (z)| ≤ |µ2 (z)|, so τ1 (z) ≥ τ2 (z).
i
Let α(z) = τ1 (z) − τ2 (z). It satisfies α(z) ≥ 0 and

∆α = |φ1 |(e2τ1 − e−2τ1 ) − |φ2 |(e2τ2 − e−2τ2 )


≤ c(e2τ1 − e2τ2 + e−2τ2 − e−2τ1 )
e2τ1 − e2τ2
= c(e2τ1 − e2τ2 + )
e2τ1 e2τ2
≤ 2c(e2τ1 − e2τ2 )
= 2ce2τ1 (1 − e−2α )
≤ 4ce2τ1 α.

Now we prove that α(z) > 0 on D. Suppose there is a point z0 ∈ D such that α(z0 ) = 0. Consider a
disk U (z0 , r ) in D with the radius r = 12 d(z0 , ∂D). Since µ1 = 2v11 and |φ1 (z)| = c > 0, the Beltrami
φ
e
coefficient µ1 (z) 6= 0 for all z ∈ D. So |µ1 (z)| has a positive lower bound on U (z0 , r ), which implies that
e2τ1 has an upper bound on U (z0 , r ). Then there is C > 0 which depends on z0 such that 4ae2τ1 < C,
so we have ∆α < Cα on U (z0 , r ). By Corollary 1 α(z) = 0 for every point z ∈ U. By considering
another point z0 ∈ U (z0 , r ) and continuing this procedure, we can obtain α(z) = 0 for all z ∈ D. So we
have α(z) = 0 for all z ∈ D, which is impossible since |φ1 | 6= |φ2 |. So we have α(z) > 0 on D, which
yields |τ1 (z)| > |τ2 (z)| for all z ∈ D.

When the Hopf differential is equal to 0, the harmonic mapping is conformal on D. So the function
γ can be defined at c = 0 with γ(0) = 0.

Lemma 10. The function γ( x ) is continuous on [0, +∞).

φi
Proof. Let φ1 and φ1 be two Hopf differentials whose Bers norms are bounded. Let µi = (i = 1, 2)
e2vi
be the corresponding Beltrami coefficient. It is known that

kµ1 − µ2 k∞ → 0 as |kφ1 k − kφ2 k| → 0.

For its proof one can see Proposition 14 of [4]. So the function γ( x ) is continuous on [0, ∞).

Lemma 11. The function γ satisfies r


4
γ(c) ≥ 1−
c
for c > 4. In particular, lim γ(c) = 1.
c→+∞

Proof. Let φ(z) be the Hopf differential of a quasiconformal harmonic mapping and µ(z) be the
Beltrami coefficient. It is known (see [4]) that if kµk∞ ≤ k < 1, the Hopf differential satisfies that

|φ| 1
≤ .
ρ2 1 − k2

|φ| 1
Thus for φ = c, we have ρ2
≤ 1− γ ( c )2
. In particular,

|φ| 1
≤ ,
4 1 − γ ( c )2
Mathematics 2020, 8, 1310 8 of 8

which yields r
4
γ(c) ≥ 1− .
c

Proof of Theorem 1. Let h(z) be a harmonic mapping whose Hopf differential is a constant c > 0.
By Lemmas 7 and 9, the function
γ ( x ) : c 7 → µ (0),

is strictly increasing from (0, +∞) to (0, 1). To prove it is a homeomorphism, we only need show γ is
continuous and surjective by Brouwer invariance of domain theorem. These are finished in Lemmas
10 and 11.

4. Conclusions
We have discussed a class of harmonic mappings whose Hopf differentials are constants. Let hc (z)
be a harmonic mapping whose Hopf differential is a constant c ∈ (0, +∞). When c is increasing from 0
to +∞, it is proved that the vaule µc (0) is strictly increasing. Since the Beltrami coefficient µc (z) of
hc (z) is radially symmetric and takes the maximum at z = 0, the quasiconformal distortion Kc of hc (z)
is also strictly increasing.

Funding: This research received no external funding.


Conflicts of Interest: The author declares no conflict of interest.

References
1. Li, P.; Tam, L.-F. Uniqueness and regularity of proper harmonic maps. Ann. Math. 1993, 137, 167–201.
[CrossRef]
2. Li, P.; Tam, L.-F. Uniqueness and regularity of proper harmonic maps II. Indiana Univ. Math. J. 1993,
42, 591–635. [CrossRef]
3. Markovic, V. Harmonic maps and the Schoen conjecture. J. Am. Math. Soc. 2017, 30, 799–817. [CrossRef]
4. Wan, T. Constant mean curvature surface, harmonic maps, and universal Teichmüller space. J. Differ. Geom.
1992, 35, 643–657. [CrossRef]
5. Wolf, M. The Teichmüller theory of harmonic maps. J. Differ. Geom. 1989, 29, 449–479. [CrossRef]
6. Tam, L.-F.; Wan, T. Harmonic diffeomorphisms between Cartan- Hadamard surfaces with prescribed
Hopf differentials. Commun. Anal. Geom. 1994, 2, 593–625. [CrossRef]
7. Yau, S.T. Harmonic functions on complete Riemannian manifolds. Commun. Pure Appl. Math. 1975, 28,
201–228. [CrossRef]
8. Cheng, S.Y.; Yau, S.T. Differential equations on Riemannian manifolds and their geometric applications.
Commun. Pure Appl. Math. 1975, 28, 333–354. [CrossRef]
9. Omori, H. Isometric immersions of Riemannian manifolds. J. Math. Soc. Jpn. 1967, 19, 205–214. [CrossRef]
10. Heinz, E. On certain nonlinear elliptic differential equations and univalent mappings. J. Anal. Math.
1956, 5, 197–272. [CrossRef]

c 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like