This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
2 and d > 1, U € L 2 ( 0 , T ; # £ A (ft)) D L°°(0,T;H* c(K) n oo L ( 7 l ) ) n J f f 1 ( 0 , r ; J f f ; 7 A 1 ( 7 l ) ) n L ° ° ( 0 , T ; ^ o o ( ^ ) ) , C/0 € ^ f a ) and / e C ( 0 , r ; i C - ; ( R ) ) , then E{uN(t) T - / 3 A / 3 T = 0, • M (c) with flat normal bundle if and only if dQx + Qx A 0 A = 0 for A € C* and 61, • • •, 6n are linearly M (c) be a non-umbilically local isometric immersion with flat normal bundle. Suppose that $\ is a solution of the system (28), /x € R \ {0}, and L is a real constant s x (2n + 1) matrix satisfying (39). Let h — LTQ^/^J^ and $A = $\D\ where D\ is the Darboux matrix given by (36). If o,-(0) ^ 0, for all j , then there exist an open neighborhood U around the origin O € M™(c) and a non-umbilically 2n—1
- U{t)) < c*N2~2r
where c* is a positive constant depending only on fj, and the norms of U, Uo and / in the mentioned spaces. The proof of the previous results can be found in Guo and Wang [26], and Wang and Guo [36]. In the last part of this section, we consider the rational spectral method for the half line. We first consider the Legendre rational spectral method for the half line, see Guo, Shen and Wang [21]. Let LJ(X) = (x + 1)~ 2 and define the space H^,(TZ+) as before. The Legendre rational function of degree I is defined by
R,(x) =
V2Lt(^). x +1
The set of Legendre rational functions is the L 2 (7?.+)-orthogonal system.
86
Let QTV = span{i?o,-fix,- • -,RN}. The L^(7^ + )-orthogonal projection + P/v : Ll1(1Z ) -+ QN is a mapping such that (PNv-v,
0 ) W = O , V0€ Q N -
In order to estimate \\PNV — v\\u, we introduce the space H^ For any non-negative integer r,
+
A(R,
).
H^ A(K+) — {v | v is measurable on 7£ + and ||w||r>U|Ji < 00} where
IHUA = E l l ( * + i)* + *flM)*. fe=0
For any real r > 0, the space iPJ A(R-+) For any v G H'iA(K+) and r'> 0, llfivt; - v\\u <
IS
defined by space interpolation.
cN-r\\v\\r,u,A-
The if^(7?.+)-orthogonal projection P/v : Hl(TZ+) -> QAT is a mapping such that for any v £ H^(1Z+), (Pkv-v,
= 0, V(^G Qjv-
In order to estimate ||P^v — ^||i, w , we introduce the space iPJ For any non-negative integer r, H^B(Tl+)
+
B(TZ
).
= {v I t> is measurable on 7£+ and ||t>||r,u>,B < +00}
where
IHIr,U,B = (X;i|(* + l)S +fc -i^«||2)i. fc=l
For any r > 0, the space H^ B(7?.+) and its norm are defined by space interpolation. For any v € HlB(Tl+) with r > 1, II^>-«III,U
lu(R-+)
= {v\v e Hl(1l+),
« ( 0 ) = 0 a n d « ( a ; ) ( x + l ) - * -> 0, as x-> 00},
Q°N = {(t> e QN I
87
and cfc(u,v) = (dxu,dx(vu))
+ v(u,v)0J.
The ^ 0 1 i W (^ + )-orthogonal projection P]fv : H^(K+) ping such that
For any v G H^{K+)
-> Q°N is a map-
n K B ( K + ) , v > | and r > 1,
Hi^>-«||i,w
0 < j < iV.
and 0 < /x < 1 < r,
l|/ATi;-«IU
l (
^).
The set of Chebyshev rational functions is the Ll1(TZ+) -orthogonal system. Let QN = span{i?o,i?i, • • • ,RN}- The L 2 (7£ + )-orthogonal projection P/v : £ w ( ^ + ) ~* QN is a mapping such that (PNv-v,4>)u=0,
V^eQjv.
In order to estimate \\PffV — v\\u, we introduce the space H£ For any non-negative integer r,
+
A(R-
)-
H£ A(R-+) = {v | v is measurable on 7?.+and |H|r,w,A < °°} where
\\v\\r\\l)lk=0
For any real r > 0, the space H* ACR>+) *S defined by space interpolation.
88
For any v G H^
and r > 0,
\\PNV - v\\u <
cN~r\\v\\rtUtA.
The i?i(7?. + )"Orthogonal projection P%, : H]1(TV') -> QAT is a mapping such that for any v € H*(H+),
(Pkv-v,4)i^=0,
V
In order to estimate \\P}fV — v||i, W) we introduce the space H„ B(R.+ ) . For any non-negative integer r, H^ B(TZ+) = {v | v is measurable on 1Z+ and ||w||r,w,j3 < +00} where
IHUfl = (Ell(* + i) 5+ *- i 5M£) i . fc=i
For any r > 0, the space H^ B(H+) interpolation. For any v 6 H^B(A) with r > 1,
and its norm are defined by space
l|P>-«||l,U
Hl(R+),
v(Q) = 0 and —^u{x)
-> 0, a s i - > 00},
71^ = {) = (dxu,dx(yuj))
+
v(u,v)u.
The # 2 (ft+)-orthogonal projection Pl/V : H^u(n+) such that ava(P1£v-v,
-> 7 ^ is a mapping
= 0, V ^ e Q V v > |± and r > 1,
l|P^>-^||i,w
89 N.
For any v € C(1l
INV(X)
G QN,
), the Chebyshev-Gauss-Radau rational interpolant
satisfying
IMtf) = «(Cf), 0 < 3 < N. For any v 6 if£ > j 4 (ft+) and 0 < n < 1 < r, l|/7vw-v|U
References 1. C. Bernardi and Y. Maday, in Handbook of Numerical Analysis, Vol.5, Techniques of Scientific Computing, ed. P. G. Ciarlet and J. L. Lions (Elsevier, Amsterdam, 1997). 2. J. P. Boyd, J. Comp. Phys., 69, 112 (1987). 3. J. P. Boyd, J. Comp. Phys., 70, 63 (1987). 4. C. Canuto, M. Y. Hussaini, A. Quarteroni and T. A. Zang, Spectral Methods in Fluid Dynamics (Springer, Berlin, 1988). 5. C. I. Christov, SIAM J. Appl. Math., 42, 1337 (1982). 6. O. Coulaud, D. Funaro and O. Kavian, Comp. Mech. in Appl. Mech and Engi., 80, 451 (1990). 7. D. Funaro and 0 . Kavian, Math. Comp., 57, 597 (1990). 8. D. Funaro, in Orthogonal Polynomials and Their Applications, ed. C. Brezinski, L. Gori and A. Ronveaux (Scientific Publishing Co., 1991). 9. D. Funaro, Polynomial Approximation of Differential Equations, (SpringerVerlag, Berlin, 1992). 10. D. Gottlieb and S. A. Orszag, Numerical Analysis of Spectral Methods: Theory and Applications, (SIAM-CBMS, Philadelphia, 1977). 11. Guo Ben-yu, Spectral Methods and Their Applications, (World Scietific, Singapore, 1998). 12. Guo Ben-yu, J. Math. Anal. Appl., 226, 180 (1998). 13. Guo Ben-yu, Math. Comp., 68, 1067 (1999). 14. Guo Ben-yu, SIAM J. Numer. Anal., 37, 621 (2000).
90 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26.
27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38.
Guo Ben-yu, J. Math. Anal. Appl, 243, 373 (2000). Guo Ben-yu, J. Comput. Math., 18, 95 (2000). Guo Ben-yu, Appl. Numer. Math., 38, 403 (2001). Guo Ben-yu, Computers and Mathematics with Applications, to appear. Guo Ben-yu and Jie Shen, Numer. Math., 86, 635 (2000). Guo Ben-yu and Jie Shen, Indiana J. of Math., 50, 181 (2001). Ben-yu Guo , Jie Shen and Zhong-qing Wang, J. of Sci. Comp., 15, 117 (2000). Ben-yu Guo , Jie Shen and Zhong-qing Wang, Int. J. Numer. Meth. Engng., to appear. Ben-yu Guo, Jie Shen and Cheng-long Xu, Advances in Comp. Math., to appear. Guo Ben-yu and Ma He-ping, J. of Comp. Math., 19, 101 (2001). Guo Ben-yu and Wang Li-lian, Advances in Comp. Math., 14, 227 (2001). Guo Ben-yu and Wang Zhong-qing, Legendre rational spectral and pseudospectral methods for nonlinear differential equations on the whole line (unpublished). Guo Ben-yu and Xu Cheng-long, RAIRO Math. Model. Numer. Anal., 34, 859 (2000). Guo Ben-yu and Xu Cheng-long, Mixed Laguerre -Legendre pseudospectral method for incompressible fluid flow in an infinite strip (unpublished). V. Iranzo and A. Falques, Comp. Meth. in Appl. Mech. and Engi., 98, 105 (1992). Ma He-ping and Guo Ben-yu, IMA J. of Numer. Anal., 2 1 , 587 (2001). Y. Maday, B. Pernaud-Thomas and H. Vandeven, Rech. Airospat., 6, 353 (1985). G. Mastroianni and G. Monegato, IMA J. of Numer. Anal, 17, 621 (1997). Jie Shen, SIAM J. Numer. Anal, 38, 1113 (2000). Wang Li-lian and Guo Ben-yu, J. Comp. Math., to appear. Wang Li-lian and Guo Ben-yu, Non-isotropic Jacobi methods for singular problems, unbounded domains and axisymmetric domains (unpublished). Wang Zhong-qing and Guo Ben-yu, A rational approximation and its applications to nonlinear differential equations on the whole line, submitted. Xu Cheng-long and Guo Ben-yu, J. Comp. Math., to appear. Xu Cheng-long and Guo Ben-yu, Advances in Comp. Math., to appear.
The Darboux Transformation and Local Isometric Immersions of Space forms Qun He Dept. of Appl. Math., Tongji Univ., Shanghai 200092, China email: [email protected] Yi-Bing Shen Dept. of Math., Zhejiang Univ., Hangzhou 310028, China email: [email protected]
Abstract By using the Darboux transformation in Soliton theory, we give the explicit construction for local isometric immersions from a space form Mn(c) into another space form M (c) via purely algebraic algorithm.
1
Introduction
The problem on isometric immersions of space forms is an interesting classical problem. There are a lot of nonexistence results in this direction ([2], [5], [8], etc.). Recently, it has been found that the integrability condition for isometric immersions of space forms, i.e., Gauss-Codazzi-Ricci equations, is equivalent to the condition of a family of connections to be flat ([6], [1]). This enable us to apply the soliton theory to the study of some problems on isometric immersions of space forms. In [6] the local isometric immersions from space forms Mn(c) into M2n(c) with flat normal bundle and linear independent curvature normals were discussed. The Darboux transformation method for the explicit expressions of such isometric immersions has been given in [9]. A soliton theory on local isometric immersions of space forms Mn(c) into M (c) with flat normal bundle and 0 ^ c / c ^ 0
91
92
was proposed in [1]. It is natural to consider the explicit constructions of general local isometric immersions from a space form M"(c) into another space form M (c). A partial work has been attempted in [3]. The purpose of this paper is to apply the Darboux transformation method to the explicit expressions of general local isometric immersions of space forms. Some preliminaries on local isometric immersions of Mn(c) into M (c) and the zero-curvature condition, which are different from [1], are given in section 2. In section 3, a solton equation for such local isometric immersions is shown, where a family of connection 1-forms including one parameter are established. In section 4, the Darboux transformations for the explicit expressions of such isometric immersions are given. This is a purely algebraic algorithm. Finally, in section 5, we give an explicit construction of such isometric immersions from a trivial (degenerated) isometric immersion via the Darboux transformation.
2
Preliminaries
Let M (c) denote an m—dimensional simply connected space form of constant sectional curvature c. Let re : M (c) -> R™+1 be the following standard isometric embedding: m
Mm(c) = {x0,xu---,xm)
<ERm+1 1 ^ 4 + ^ ,4=1
*?"(£) = {(x0,Xl,---,xm)
e RT* \Y,*A-*l
-
= ^}
= --k)
A=l
W(Q)
= {(x0,x1,---,xm)£Rm+1
for
g
>0,
c
^
c<0,
°
|x0=0}.
Let Mn(c) be an n—dimensional space form of constant curvature c, and U C Mn(c) an open neighborhood. Consider a locally isometric immersion V? : U -¥ M (c) where m > n and, without loss of generality, c = ± 1 or 0. We shall make use of the following convention on the ranges of indices unless otherwise stated: i,j,k,---
= l,---,n;
a,/3, • • • = 1, • • • ,m - n.
Then the composition map r = r e o ip : U -4 R™+1 is a local isometric immersion into i2™+1. Set J e =
( o
7°)'
sOc(m + l) = {X esl{m + l,R)\XJ-c
+ JcXT = 0}.
93 Denote by SOc(m +1) the Lie group, of which the Lie algebra is so e (m +1). Consider a framing field \P = (eo, e\, • • •, em) : U —• SOc(m+l) in R™+1 so that r = J j e 0 , {e;} are tangent to Mn(c), and {e n + Q } are normal to Mn{c) in M (c). Clearly, eo is normal to M (c) for c ^ 0. Let H = * _ 1 dvf be the pull back of the Maurer-Cartan form of SOc(m + 1) by * , which is an sOc{m + 1)—valued 1-form. We then have
where s =
SL-7 \ f(o)-w,
' w £ ' yQ _pT v j
(1)
where 8 = (61, • • • ,6n)T are dual fields of {e;}, u — (uy) is the Levi-Civita connection 1-forms of Mn(c), 0 — (cjj>n+0() is the second fundamental form of the isometric immersion
d0 + PAri + u/\P = O,
(2)
T
A7 + 77 A 77 - /3 A /3 = 0, 9T A /3 = 0. Since M n (c) has constant curvatures c, then duj + u)Au = c6A0T.
(3)
It follows from (2) 2 and (3) that PAf
+ (c - c)fl A 6T = 0.
(4)
If the normal bundle of (p is flat, then dr] + 77 A 77 = /3 T A /? = 0.
(5)
Set , . e = sgn(c), /_ > • e = sgn(c-c),
f 1 J /— 1 1/=^ /p
for c = 0,
K=
r
for c = c, _
w
94
Then (4) can be rewritten as /3 A /? T + ev29 A 0T = 0.
(7)
Consider the Lie algebra soex(m + 2) = {X G sl(m + 2)\XJ + JXT = 0},J The Lie group SOex {m + 2) corresponding to soex (m + 2) is SOex{m + 2) = {A € SL{m + 2)\AJAT
= J}.
We now define a family of soex(m + 2, C)-valued 1-forms parameterized by A € C* = C \ {0} as follows / 0 0A =
-enOT
K0
0 \ 0
W
-A£T -\evQT
0 A£
0 \ \V8
(8)
0 0 /
T?
0
By (2)~(8), we have the following Lemma 2.1 There exists a local isometric immersion
(9)
independent.
When the isometric immersion tp has flat normal bundle, the second fundamental of (p can be simultaneously diagonalized. In such a case, we can choose the tangent frame fields je*} to Mn(c) such that Wj, n+Q = bia6l. Moreover, we can choose a parallel normal frame fields {eQ} so that rj = 0. On putting k
we have from (2), (4) and (7) ^2 kabja + EV2 = 0,
(i ^ j)
a
(bia - bja)T^ = {bia - bka)Tjik, ej{bia) = (bja - bicJFij.
(i,j,k
(i ^ j)
#
(10)
95 Set y = (tv-^)r,
fi = (6te),
n
Js=^~n
B = (B V), Then we see from (8) that BJgBT
°).
(11)
= diag(/9i, • • • pn), where
Pi = £ ^
+ £ V
"
(12)
L e m m a 2.2 Let <^ : Mn(c) D U -> M (c) oe a locally isometric immersion with flat normal bundle. Assume that Pi ^ 0 for all i where pi are smooth functions defined by (12). Then there exist a line of curvature coordinates (xi) on U such that the first and second fundamental forms of (f can be given by
i
II = ^2oJfbiadxfen+a.
(13)
Proof. Since Pi ^ 0 then we can write pi = ±(a;) follows from (10) and (12) ejiai) = aiTlj,
2
with a; > 0. It
(t#j).
(14)
For any point x e U, if we choose e n + i at x so that bn(x) ^ 0, 612(2;) = • • • = bi,m-n(x) = 0, then it follows from (10)i and (12) that M z ) = - 7 - T T (*? £ 1 )' &ii(*) + e>2 = ^ 0 . (15) On (a;) By taking a — j — 1 in (10)2, we see from (15) that T^(x) = 0. Since x is arbitrary, then it follows that T*j = 0 for i, k, 1 distinct. We know that the components T^ of Wjj are independent of the choice of the fields of normal frames. Thus, we have T*- = 0 for i,j, k distinct. Hence, by using (14) and the skew-symmetry, we conclude that Wi. =
e
e
jM6i _ Oj
-iMej.
(16)
Oj
As the same as in [4], it is easy from (16) to see that there exist a line of curvature coordinates (x») on U so that d/dxi = a;ej, 9l = aidxi. Thus, the first and second fundamental forms of ip are given by (13).
96
In the case that e > 0, we see from (10) that rankZ? = n, which implies that m — n > n — 1 and Pi > 0 for all i, so that the conclusion of Lemma 2.2 holds. In the case that e = 0, if m — n > n and the normal bundle of
= Js. So, by Lemma 2.2, we have immedi-
Proposition 2.3 Let ip : Mn(c) D U -)• "M™^), c ^ c, be a local isometric immersion with the flat normal bundle. If
I = ^2a2dx2, i
II = 22 aiaiadx2en+a,
(17)
and A = (atj) : Rn -> SOe(n) = {X € SL(n)\XJeXT
= Js),
(18)
where a.in = i/ai.
3
Non-umbilically isometric immersions of space forms 2n—1
Consider a local isometric immersion ip : Mn(c) -> M Under the same hypothesis as in Proposition 2.3, we set fij = ^r(i^j), a j
b = K(ai,---,an)T,
f« = 0,
A1 = (Aia),
(c) with c ^ c. F =
S = dia,g(dxi,---,dxn).
(fij), (19)
97
We then see that q = K-lvb
= AEn <E 5 n - 1 (£),
F € gl(n)* = {Y = (yij) e gl(n) | yu = 0},
where En = diag(0,• • • ,0,1). Choose parallel frame fields in the normal n-l
bundle so that n — 0. Thus, S of (1) and 0 A of (8) are reduced as / 0 S = K-X5b
-cK~1bT6 OJ
0 \ SAi ,
V0
-A{6
0 /
0A
/ 0 -ebTS = \ 5b u \ 0 -\JSAT5
Lemma 3.1 Let h : Rn -* R2n+1 h as a row vector 1 n
(20)
0 \ XSA . 0 /
(21)
satisfy the equation dh = h@i.
Write
n—1 1
(Z v
C
C)
where C satisfies that C_(0) = K _ 1 I / ^ ( 0 ) . Then h = satisfies the equation dh = hE.
(KT^TJ.C)
: Rn -> -R2"
Proof. Since /i satisfies d/i = Ji0i, then we have from (21) d£ = r)6b, dr] = T]u}- CA{5 - (ef + £K~lvC,)bT8, dC,=r]6Al, dC, = K~lun8b — K~1ud^. By the last equation and the initial condition, we see that (, — K _ 1 Z/£. It follows that ef + e/c _ 1 K = K ~ 2 ( K 2 £ + ev2)^ - CK~2^. Hence, we have h = ( K - 1 £ , 7 J , £ ) satisfies dh = h!E. For simplicity, we write a p x (2n + 1) matrix M as a row matrix I
n n MW
(MM
M^).
Particularly, we write a (2n + 1) x (2rc + 1) matrix M as a block matrix I
M(n)
n
M(n)
n
^(13)
x
^(21)
^(22)
^(23)
n
M(ZX)
M{Z2)
M(33)
n
•
98
By using the gauge transformation ©A = HQxH'1
- dHH~\
wheretf = (
7
"+ 2
^ J € SOex{2n + 1), (22)
we obtain ex=\
/ 0 -ebTS Sb u
0 \ XS ,
\ 0
0 J
-XJeS
(23)
where w = 5F - FT5,
0 = JgSFTJe - FS.
(24)
It is easy to see that dQ\ + ©A A6A = 0 if and only if d&\ + Q\ A 0 A = 0, which is equivalent to that (F, A) satisfies the following system of PDE: dAi = -i?Ai, db = - 0 6 , dw + w A w - e<56 A 6 T 5 = 0,
(25)
i.e., the Gauss-Codazzi-Ricci equations for the isometric immersion
0 0 0 0 0 -J£5
0 \ / 0 6 \,v = 0
0 0 V b F
0 /
-ebTJ-E -FTJe-
,
=
0
In+1 0
0 -In (26)
we have ©A = a\+ b = vW,
[a,v], F = vW.
(27)
With respect to a, the Lie algebra Q = soex(2n +1) has the Cartan decompositions Q = V ®K. Let Qa = {y G a|[o,»] = 0},
Qi = {z € Q\tr{zy) = 0 for y € Ga},
KaQ = {X{\) G /\Q\
X(-X)}.
Clearly, a is P-valued 1-form, v : U -> g^ f l ? is a smooth map. Thus, ©A is a Aag—valued 1-form. Consider the system f d$A = *A(OA + [a,«]) = $A©A, 1
$A(0) = / 2 „ + I ,
MR1 {2
*>
99 of which the integrabihty condition is (25). For the solution $A to (28), we have A=(je-*i*3)Jg)TA(0). (29) Set $ A = Q i t f _ 1 ( 0 ) $ A # , where 0 0
K~lV 0
Ai(0)
0
/ 1 0 Qi = 0 /„
V0
0
Then $ A satisfies d# A
= $A9A.
Write $i = (r,ei,---,e2„),
f =-r,
* = (f,ei,-• • , e 2 n _ i ) .
K
By a straightforward calculation, one can see that e 2 „(0) = UK~1T(0). it follows from Lemma 3.1 that * satisfies the following system
d* = * ~ ,
SO,
*(0) =
where S is defined in (20). Set K
* = Q2
0 In
Then $ : [ / ' - > 5 0 a ( 2 n ) satisfies the system (1). Hence, we obtain r = j | Q 2 f = JlQiQiH-HO^K^H^
= QQ™,
(30)
where Q =
/ c2
0
0
1
V 0
c(l - C C ) K - 2 6 ( 0 ) T J f
/c- /,,
\
0
/t-1A1(0)TJE-
0
(31)
/
is a constant 2n x (2n +1) matrix. Summing up and combining Proposition 2.3, we have proved the following Theorem 3.2
Let U C Mn(c) be a simply-connected domain around the 2 n—1
origin x — 0, and (p : U —• M (c), c ^ c, a non-umbilically local isometric immersion with flat normal bundle. Then there exists a smooth
100
map (F,q) : U -» gl(n)» x Sn 1 (e) such that ®\ defined by (23) is a flat connection and the system (28) has a unique solution $\ satisfying r = r e O ^ = Q$< 1) .
(32)
Conversely, for a map (F,q) : Rn —> gl(ri)* x 5 n - 1 ( e ) , if (28) has a unique solution $\, then there exists a smooth map A = (ay) : U -¥ SOg(n) such that q = AEn. Moreover, if U = {x G Rn\a,i(x) ^ Oforalh'} is not empty, 2n—1
then there exists a non-umbilically isometric immersion
4
Darboux transformation
We now consider the Darboux transformation for solutions of the system (28). Since Q\= a\ + [a,u] is a l\aQ—valued 1-form, then §\ satisfies the following G/K-reality condition (cf. [7]): f(\)Jf(\y
= J,
/(A) = /(A),
af{X)a = f{-\).
(33)
Let Ooo be an open neighborhood around oo in CU{oo} = S2, and let G™ = { / : Ooo -> GL(N, C)|/is a holomorphic rational fraction satisfying (33)i and /(oo) = IN}, {G™)a = {/(A) € G™|/(A) satisfies (33)}. A map 7r : V —>• CN is called a J—Hermitian projection if 7r2 = IT,
Jit* = nJ.
Clearly, n' = I — n is also a J—Hermitian projection if 7r is one. Thus, a simple element of (7™ is of the form [7]: ,,.
.
X— a
5a,*(A) = TT' + T
a —a
T7T - / -
.,.
-TT
34
A—a A—a for a 6 C* = C\{0}. Let r be a diagonal complex matrix satisfying r 2 = a. A direct computation yields the following Lemma 4.1
Let TTO be a J—Hermitian projection in CN ^0 =
TTO)
CTroCTro =
7ro
satisfying
101
If-jr = T
1
TT0T,
then /(A) = ga^9-a^G
e (G™)CT for a €
y/^R.
Let $ A be a solution to (28), L a constant complex s x N matrix satisfying LJaLT = 0, det(LJLT) ^ 0. (35) Set TTO = JLT(LJLT)-1L, /i = Lr $ a ,
7r = T_1jroT,
Ti1 = Jh* (hJh*)-*/»,
*A = 3a,7r*A5a,7Ti, & = Lra^!-a, ff2 = Jh* where a = V—lfi for /i 6 i? \ {0}. Then, we may write the following Darboux matrix
(hJh*)-1^,
(36)
Thus, we have following [3] Theorem 4.2 Let 3>A &e a solution of the system (28), and L a real constant sxN matrix. Seth = L r ^ y r y ^ forn € i?\{0}. Then there is an open neighborhood U around the origin 0 such that on U, $ A = D\(0)~1$\D\ satisfies the system (28) with v = v + (di)g± : [/ -> ^ fl P , namely, a\+[a,v] is a Aa—valued 1-form, where D\ andd\ are defined by (36) and (37), respectively. In the following, we take N = 2n + l. By Theorem 3.2 and Theorem 4.2, it is sufficient to find the Darboux matrix (36) preserving q(x) e Sn~1(e). By (25), we can see easily the following Lemma 4.3 Let 4>A be a solution of the system (28), L a complex constant s x (2n + 1) matrix. Assume that Ao 6 C and h = L $ A 0 = (£,??, ()• Then d((b - A0£) = 0. Let ft G R \ {0}, h = Lr^^z^^
= (£, n, y/^lC,)
satisfying
d£ = nSb, dn = -e$,bT8 + TJUJ + d£ = nrjS + 0?.
tfJgS,
(38)
102
By Theorem 4.2 and Lemma 4.3, if we choose L such that L satisfies (35) and L^b(0)T - AiL<1> = 0, (39) then there exists an open neighborhood U around the origin 0 such that h = (£, r), >/—TC) satisfies hJah* = stfT + nnT - C, JsCT = 0, det(/iJ/i*) ^ 0,
C&T - A*f = 0,
(40)
in U. Thus, (36) can be written as
£>A = / - T ^ 7 - T (
peewit WTAif AJK T Ai£
veZTAiV wTAiU AJ^Anj
-Ae£ T A 2 C -AT7TA2C
/^CTA2<
),
(41)
where Ai = (A + Aip)_1,
A2 = ( A - / x p ) - 1 ,
A = i/ijft* = CJe-CT = e « T + W,
(42)
p = -/»Wfc* = C ' J r f C T - e ^ r - i / , U r . Here/i'_=f = (C',^V=TC'). Set $ A = $ A A \ - It is easy from Theorem 4.2 to see that $ A satisfies the system (28) with v - v + (di) 6 x. Thus, by (27) and (29), we have F =(«(»)) \
b=v ^
=F-2M^CTA17?)off,
/ Off
= b _ 2//J f C T Aie =
K-HOI,
• • •, a „ ) T ,
(43)
i = ( j f ^ 3 3 ) J f ) T A(0) = A - 2JfCTAiCv4. Since £b = /x£, then q = K^vb = AEn. Noting that A 6 SOg(n), we see that q e Sn~1{e). Hence, we have the following 2n—l
Theorem 3.4 Let
local isometric immersion (p : U -> M (c) such that f = re ° (p can be expressed explicitly by f = QD^iO)*™ = QD^WQtD™, (44)
103 where Q is a constant matrix denned by (31) with A(0). Remark 1. The above process of the Darboux transformation is purely algebraic. Hence, starting from a special solution A to (28) for which the corresponding r = Q$[ ' may be degenerated, we can repeat the processes via the purely algebraic algorithm and obtain a sequence of solutions to (28): $ A -> $ A - • $ A -> • • •, from which we obtain a sequence of local 2n—1 isometric immersions from Mn(c) to M (c). Remark 2 The method here may be used to study local isometric immersions from Riemannian products of space forms into space forms. We are going to discuss these problems in a forthcoming paper.
5
The construction of local isometric immersions derived from a trivial solution
By take a trivial solution of (25) as F = 0,
A = In,
6 =
(0,---,0,KO,
We solve (28) to get /
fc];'1
0
0
Xi
0 "X*
$>
n
0
\J
•••0
• •• u
Yn-i Xn
-Yt
-y„-i
0
Xn-i sxYn
0
0 Ay Xr x 0
Xi
•••
e^$A! ' m \
C
0
Yi
X„-i
o ,ro £9 e*l X
-f£Y "X n 0
n
0
\
•••0
0 $ in,tit
where
•i' 1 = ^ ( « a * « + e _ A 2 ) '
*rm =
^xn+^),
,
(45)
104 K\
"A.
Xa = cos(Aa;a),
Xn = cos(xz n ),
Ya = sin(Ax a ),
Yn = sin(xx„).
Choose fj, € R and i = (lo, h, • • •, hn) 6 R2n+1 5 Z '"+« + a 2 n = £ ' ? + ei o ^ 0, a
such that
«Z2n - A*i//0 = 0.
(46)
j
It is easily seen that h =
/T$V^TM
= (f,
JJ,
V^IC) = (€.i7i»---.»?n,V/::TCi,---,"v/=:TCn)
satisfies (38) and (40). From (41), (42) and (43) we know that
"i"" ( O T (c + ^ - ^ , - ^ „ f «*)*, A = £
C* + eCn,
A = J n - I JeCTC-
(47)
a
By Theorem 4.4, We can use directly the formula f = Q£>j"1(0)$-Di in a neighborhood U of the origin. Such f is nondegenerate only if there exists a point x £ U such that bj(x) ^ 0 for all j . For this aim, we need only to choose suitably I such that ln+j^0,
£ £
+
a + (£-2)JL^0.
(48)
a
It may guarantee that bj(0) ^ 0 for all j . Then there exists an open neighborhood U of O such that f is nondegenerate in U, which implies that -
2n—1
there is a local isometric immersion ip : U -> M (c) so that f = r g o (p. Moreover, by using $Ai we can obtain a new solution $ A of (28). Continuing this process, a series of local isometric immersions are obtained via a purely algebraic algorithm. For simplicity, in the case that 7 = y/en2 — EKV~X 6 R, we may take I = (lo,h,m •-,'n-i>7' / K ~ '0>'1 j • • • Jn), where ln = IIVK^IQ, IJ ^ 0 for all j , and £ ) a ^ + (^ _ 2)/£ ^ 0. It is clear that (46) and (48) are satisfied. Then we have from (45), (47) and (43) £ = l0e~^
= -Cn,
105 'la — l a c
— sco
6Q = - | W n e ^ ° + 7 1 " ,
(49)
a
In the case that 7 € ^/^Ti?, we may take Z = ( Z o , i l , - - - , J n - 2 ) 0 , 0 , Z i , - - - , Z n _ 2 , V—l7/i"~
In,In),
where Zn = HVK~1IQ, Z Q _I ^ 0 for all a, and ^ I " ^ 2 l? + (e-2fj?v2K~2)ll ^ 0. The remainder is similar to the above. Example n — 2. c = e = /i = v = K = l. c — 0, E = —1, x = ^/T^A2, 7 = v^=2. By taking I = (Zo,0,0,-\/2Zo,Zo) with Z0 ^ 0, we have from (45), (47) and (43) £ = C2 = Zo cos V2x2, t]2 =-losinV2x2,
m = -V2lo sinh xi, Ci = - v ^ / o c o s h x i ,
/
f=-^f
v 2 cos v2a;2 (sinh Xi cos xi + cosh Xi sin xi) a;2A+^sin2v^x2 \ \/2A + v^cosv^2x2(shxi sinxi - chxi cos^i)
7 _ 1 ( -2ch 2 xi - cos2 X2 ~ A V -2v / 2cha;i cos y/2x2
2-\/2cha;i cos y/2x2 2ch 2 xi + cos2 y/2x2
where A = 2ch2a;i - cos2 v2x2.
Thus, f is a piece of non-umbilical surface with constant Gauss curvature 1 in R3. Acknowledgments Project supported by the National Natural Science Foundation of China, the Scientific Foundation of the National Education Department of China, and the Natural Science Foundation of Zhejiang Province.
106
References [1] D. Ferus and F. Pedit, Math. Ann. 305, 329(1996). [2] D. Hilbert, Trans. AMS2, 87(1901). [3] Q. He and Y. B. Shen, Explicit construction for local isometric immersions of space forms, (preprint). [4] J. D. Moore, Pacific J. Math. 40, 157(1972). [5] F. Pedit, Comm. Math. Helv. 63, 672(1988). [6] C. L. Terng, J. Diff. Geom. 45, 407(1997). [7] C. L. Terng and K. Uhlenbeck, Comm. Pure and Appl. Math.53, 1(2000). [8] F. Xavier, Comm. Math. Helv. 60, 280(1985). [9] Z. X. Zhou, Inverse Problems 14 , 1353(1998).
On the Nirenberg Problem Ji, Min (institute of mathematics,
Academia Sinica, Beijing 100080, China)
[email protected]
Abstract. In this paper we strengthen the Kazdan-Warner necessary condition so as to be necessary and sufficient for the Nirenberg problem. Enlightened by this condition, we establish a degree theory which improves some important results previous and gives some results completely new.
§1.
Introduction
We begin with the prescribed scalar curvature problem on a general Riemannian manifold (M, go) which is compact and of dimension two. Given a continuous function R on M, whether can R be the scalar curvature of some metric g which is pointwise conformal to the original metric
- Agou
+ Ro-Reu
= 0,
on
M,
where i?o is the scalar curvature of the metric go • It is related to topology of the manifold M. Integrating the equation (1.1) and using Gauss-Bonnet theorem, we see [ ReudA = 4irx(M) JM where dA is the volume element of (M, g0) and x(M) is the Euler-Poincare characteristic. Clearly (1.2) requires different conditions of R for different signs of x(M). (1.2)
* Supported in part by the NNSF of China(Grant No. 19725102) and the 973 project of China.
107
108
In the case x(M) < 0, ones have had a good knowledge for the solvability. For example, it is known that (1.1) always admits a solution if R < 0 (cf. [19]). In the case x(-^0 = 0, the problem was solved by M.S. Berger and Kazdan-Warner, who gave the following necessary and sufficient condition(Theorem 6.1 in [19]) B - K - W Theorem. Suppose x(-W) = 0. The equation (1.1) is solvable if and only if either (1) R = 0, or (2) R changes sign and satisfies E 0 n BKW jt 0, where So =
se
t of the solutions to A 9o u — Ro,
and BKW = {u&H1(M):
f
R eudA < 0}.
JM
In the case x(-W) > 0, there are only two possibilities that M = S2 with x(-W) = 2 and M — RP2 , the real projective space, with x(M) = 1. From (1.2) we see easily a necessary condition that R is positive somewhere, which is proved also sufficient for the latter case(cf. [1], [22]). Thus, the former case M = S2 is only remaining and also difficult. If (M, g0) is the standard 2-sphere (S 2 , go), the problem is called the Nirenberg problem. Now Let S2 = {x = (xi,x2,x3)
e R3 : x\ + x\ + x\ = 1}
with the natural metric go — dx\ + dx\ + dx\. Let y a n d A be the gradient and the Laplacian operator on (S2,go) respectively. Actually the Nirenberg problem is to determine which smooth function R > 0 can be the scalar curvature of a conformal metric g — eugo (see [20]), i.e. to determine function R > 0 such that the following PDE: (1.3)
-Au
+ 2-Reu
= 0,
on
S2
109
has a solution. Below we shall always assume R > 0 and discuss the solvability of the equation (1.3). In section 2, we recall a necessary condition, say, Kazdan-Warner condition, and give a necessary and sufficient condition (Theorem 2.1). Section 3 is about the existence of solutions of (1.3), including a degree theoryTheorem 3.1 or Theorem 3.2, which is shown to improve some important results previous (Corollary 3.1-Corollary 3.3). The main steps for proofs of these two theorems will be given in section 5. And we shall collect in section 4 some basic inequalities and estimates previous that plays a fundamental role in the study of the Nirenberg problem.
§2. K a z d a n - W a r n e r condition The solvability of (1.3) can be reduced to a variational problem: a critical point of the functional (2.0)
J{u) = \f | v « | 2 - 8 7 r l o g / ReudA, 2 2 Js Js2
Vu €
ff1^2),
yields a solution to (1.3). This functional is bounded from below, however it has no minimum except R =const. and is lacking of compactness. For the solvability, various kinds of sufficient conditions are given, which we are going to talk about in next section in more detail. First of all, it is known that it is not always solvable and there is a necessary condition, Kazdan-Warner condition, which says KW^0 where (2.1)
KW—
{ueH^S2)
: / sjR- v ^ i e" dA = 0, i = 1,2,3}. Js2 This gives rise to many examples for which (1.3) has no solution. For example, if the function R is taken to be R = x\ + c with c > 1 a constant, / 2 \jR-sjxxeu Js
dA=
I 2 | ya:i|2 Js
eudA>0
110
for any u G H1(S2), then the set KW is empty and the equation (1.3) has no solution by Kazdan-Warner condition. Precisely speaking, they proved ([19]) the following K - W Theorem. If u solves (1.3), then u G KW. Proof. For u G H1, set Vk = y u ' V 1 * (k — 1,2,3). A direct calculus yields / y u • \7vkdA Js2 / V u • V ( V U • \/xk)dA
-
V(l V "I 2 ) ' S/XkdA - / 2 | v Js2 Js A / 2 | y u | 2 ( - A i t -2xfc)d^ Js 'S2 0
=
h
= =
u\2xkdA
and /
ReuvkdA
2
h =
/
=
-
=
2
R v (e u ) • yx fc cL4 / y i ? • yz f c e u cL4 - / i? e u A a;fccL4 .As2 Vs=
Vs2
R euxkdA
- / 2 y i ? • yz fc e u cL4 ^s
since — Axk = 2 Xk
k = 1,2,3.
It follows by the definition of J that for k = 1,2,3, < dJ(u), ufc > =
/ y u • yufcdA + 2 / ufccL4 Js2 Js2
/ y u • \7xkdA Js2 + -Mj Reu
[ s
J 2
j I Reu
\/R-\/xkeudA
^^— / R / Reu Js2
/ R Js2
euxkdA
euvkdA
111
and < dJ(u), xk > = / V " • \jxkdA Js2 since
L is2
-r /
Reu
/ R Js2
euxkdA
xkdA — 0.
Thus we get (2.2)
< dJ(u), vk>=2<
dJ{u), xk > +—
/ u E>„« Re
/
V-R • V ^ e u dA
JS2 Js2
for any u £ Hl. Now let u be a solution, then < dJ(u), vk >= 0; < dJ(u), xk >— 0, which together with (2.2) yield that
L is2
V # • S?xkeudA = 0
A; = 1,2,3.
This means u £ KW, finishing the proof.
•
Whether is Kazdan-Warner condition sufficient? In 1987 BourguignonEzin [2] found a slightly more general condition than Kazdan-Warner condition obstructing the solvability of (1.3) and gave an example of R that satisfies Kazdan-Wamer condition, but not theirs. In 1995 Chen-Li ([10], [11]) obtained some new nonexistence results where R is axi-symmetric and satisfies the Kazdan-Warner condition. All of these illustrates that the Kazdan-Warner condition is not sufficient. Then a problem is how far Kazdan-Warner condition being from sufficiency. Below we suitably strengthen Kazdan-Warner condition, so as to be both necessary and sufficient. Let Ho be the Hilbert space flo^uefl^2):
/ u = Q}. Js2
112
Denote B = {x £ R3 : x\ + x\ + x\ < 1}. Like Chen-Ding [8] and the others, we introduce the center-of-mass map P : H1 ->• B by (2.3)
P(u)=
[ xeu^dx Js2
/ [ e<xUx, I Js2
VuGH1.
For a £ B, set (2.4)
Ma = {ueH0:
P(u) = a}.
It is not difficult to prove that Ma is a submanifold of the natural Finsler structure in H0, by showing that the differential operators {a!Pi(u)}i
Ka = {u£Ma:
dJa(u) = 0}.
Finally let
(2.6)
S := (J
Ka.
Theorem 2.1. The equation (1.3) is solvable if and only if (2.7)
SflKW#0
where the sets KW and S are defined by (2.1) and (2.6) respectively. Proof. Let u be a solution. By K-W theorem, u €KW. And by GaussBonnet theorem, /
Js2
ReudA = 87r, then u is a critical point of J. It is
naturally critical for the restriction Ja with o = P(u), then u e Ka C S. Thus u £ S fl KW. The necessity is shown. For sufficiency, let u £ SnKW. Then u £ Ka for some a £ B, a critical point of the functional J a . Below it will be shown from u £KW that u must be also critical for J. Hence u + c solves (1.3) for some constant c. The solvability is obtained. Indeed we first characterize the tangent space of M n at u, i.e., Tu(Ma). By P(u) = a, / eu(x - a) vdA < dP(u), v > = ^ — / eudA Js2
W £ Ho,
113
we then have (2.8)
Tu(Ma)
= {veH0:
I eu{x - a) vdA = 0}. Js2
Denote (2.9)
T
H = /2 Js
eU x
i i ~ a,i){xj - a,j)
(i,j = 1,2,3),
then T := (^3)3x3 is a positively definite matrix. We denote T For v eH^S2), set
l
= (T 8J ).
v := —- I vdA\ ±* Jap 's2 (2.10)
ai=
2
eu{xi-ai)v
(i = 1,2,3);
Js
pi = ^ajTi'
(2.11)
(* = 1.2,3);
and set {; : = (v -v)
(2.12)
We claim that v 6 Tu(Ma).
-y^^PiXj.
Indeed, / vdA = - ^ Pi / %idA = 0, i.e. Js2 ~[ Js2
v € Ho, and for j = 1,2,3, / eu(xj — a,j)v = / eu(a;j — aj) (u — v) — / J $ 3 is2 Js 2 [ j=1 =
/ 2 eu(a;j - a , > ~Y]Pi Vs ~J
Js
2
eu(a;i - oi)(x:(- - a.,)
3
= ay-£>!!,• = 0 »=i
by using P(u) = a and (2.9)-(2.11), thus u € Tu(Ma) in terms of (2.8), and the assertion is obtained. Now from dJa(u) = 0 we have
VwGff 1
114
where v is defined by (2.12). Choose v = vk = V w " \7xk {k = 1,2,3) and aki is determined by (2.10) with v = vk, i.e., (2.13)
aki = /
eu(Xi - en) vk dA
Vfc.i = 1,2,3,
and 3
Pki = Y,ai
(1.14)
(*,» = 1,2,3),
then we have 3
vk:=(vk-vk)-^2PkiXi
eTu{Ma)
(k = 1,2,3).
2=1
It follows that (2.15)
< dJ(u), vk >= 0
i.e. 3
(2.16)
vk-vk>=Y^Pki
Xi>
Vfc = 1,2,3.
»=i
According to the identity (2.2) (in the proof of K-W Theorem) and the fact that < dJ(u), c > = 0 for any constant c, we see the left hand side of (2.16) equals 2 < dJ(u), xk > H— / V-R • \ Reu Js2
\/xkeudA,
equaling 2 < dJ(u), xk > by u £ KW. Thus (2.16) becomes 3
(2.17)
2
xk >=YjPki
Xi
>,
t=i
where (3ki is defined in terms of aki by (2.13) and (2.14).
Vfc=l,2,3
115
Carefully computing aki in (2.13), we have that for k,i — 1,2,3, Q-ki =
eu(%i - flj) V w •
/
S/xkdA
(xi - a) V (eu) • S7xkdA
s2 / /
eu v Xi- \/xkdA
- /
eu\7Xi-yxkdA
+2
eu(xi - a*) A eu{xt-
e" V xi- VxkdA + 2 euVXi5
xkdA,
a,i)xkdA
eu{xi - a,i)(xk - afc)d,4
VxkdA + 2Tki
2
by making use of — A xk = 2 xk and P(u) = a and (2.9). Denote Gki=
e" V Xi • \jxkdA,
k,i = 1,2,3,
and denote G := (Gki)3x3, which is a positively definite matrix. Then we have the expression: (aki) = — G + 2T, and (Pki) — (aki) T - 1 = -GT-X + 27 by (2.14). Substituting it into (2.17) and noticing that det \G\ det \T\~l > 0, we obtain
VuGH 1 ,
with u is defined by (2.12), imply that < dJ(u), v>=0 Thus dJ(u) = 0. The sufficiency is shown. The proof is finished.
v G H1
Q
Remark 2.2. It is good that the restricted functional Ja, for each parameter a £ B, is compact, satisfying the Palais-Smale condition. Hence there is no any difficulty in analysis to determine the critical set Ka and so S. At least, Ja has a minimum ua (Va & B), and S contains all of these minima. By carefully investigating these minima and the set KW, we shall arrive in the next section at a degree theory for the solvability.
116
The point of Theorem 2.1 is to turn the problem for the bad functional J (lacking compactness) into that for the good functionals Ja (being compact) (a e B) through Kazdan-Warner set KW. In other words, the set KW bears the responsibility for the lacking of compactness of the problem. Clearly, an investigation of the set KW is also helpful in understanding the problem. Roughly speaking, the bigger KW is, the more possible the equation (1.3) admits a solution. For example, when R is a constant, the set KW is biggest, being the whole H1, then E n KW=E is not empty, leading to the solvability, a well known fact. Another example is the case where R is even. In this case, we observe that {even functions} CKW. Indeed, because of the symmetry, it suffices to show (2.18)
/ v # - VZ3 e" dA = 0 Js2
for even functions u. Let (8,(/>), —TT/2 < 8 < ir/2, —ir <
s/R-\/x3eu
0) = u(-0,
U{6,
is an odd function of 0, then dA=
f J—x/2
cos2 6d0 f
i?0(0,>)e u(M) # = 0.
J-iT
This verifies (2.18) Thus, the set S n KW contains S D {even functions}. Moreover, it is clear that E n {even functions} ^ 0 since R is even. We see E n KW / 0, and the equation (1.3) is solvable by Theorem 2.1. This is just the famous result due to Moser ([21]).
§3. Existence of solutions In recent almost thirty years, there have been a lot of works about the existence of solutions to (1.3) and various kinds of sufficient conditions have been given.
117
The pioneer work is due to Moser who proved the existence in 1973 for R being even(as we mentioned in last section) by minimizing the functional J in the space of even functions. Later Hong([15]), Xu-Yang([25]) considered R as a axi-symmetric function. And Chen-Ding([8]), Ji([17]) studied the case where R is general symmetric, say G-symmetric, R(Gx) = R(x), Va; £ S2, with G an orthogonal transformation group on S 2 . In these special cases, some other conditions are imposed for the solvability. When R is not necessarily symmetric, there are also many studies([3]-[7], [9], [12], [14], [16]). Since the functional J has no minima in general, people made efforts to look for minimax type of solutions, such as, by establishing a mountainpass-lemma, Morse theory, and other minimax schemes. The following result is due to Chang-Yang ([5],[6]): Chang-Yang Theorem. Suppose R has only isolated non-degenerate critical points and in addition satisfies (3.1)
| v -Rl + I A H| ^ 0.
If £y6S_(-i)ind(^i, then the equation (1.3) has a solution, where 5_ := { y e S 2 : vR(y)
= 0, AR(y) < 0}.
This result was proved, as a main result, first in [5] by using a delicate minimax procedure and analyzing where the minimax sequence fails to converge, later proved in [4] and [14] by establishing a Morse theory for the functional J. Later in [3] and [7], the existence of solutions to (1.3) was shown when a map G, associated to the function R, has non-zero degree. Here the map G was defined using the action of the conformal group of S2, and has an integral expression. They considered the following set of conformal transformations of S2. Given P e S2, t > 1, using y as the stereographic projection from S2 — {P} (where P is the north pole) to the equatorial plane 2/i,2/2- Let (j)pj be the conformal map of S2 given by (f)p,t(y) = ty. The totality of all such transformations comprise a set which is diffeomorphic to
118
the unit ball B, with the identity transformation identified with the origin in B and (j>ptt o ( "7 )P — p G B in general. They constructed the map G : B -> R3 by setting (3.2)
G(P,t)=
[
(Ro(f>Pt).xdx.
It was motivated by Kazdan-Warner condition since the condition than expresses the fact that the variation of J vanishes along the conformal transformations. They proved that if (3.1) holds, and if deg(G,B,0) ^ 0, then the equation (1.3) has a solution. As an application of it, Chang-Yang Theorem is derived again in [3]. Enlightened by Theorem 2.1, a necessary and sufficient condition to the solvability of (1.3), below we are going to investigate the sets KW and S. In fact, we shall not consider the whole X, but only a subset of it, the set of minima of the functional Ja for all a G B. Notice the fact that (3.3)
J(u + c) = J(u),
P(u + c) = P{u)
for any constant c. For a G B, we set (3.4)
Ma = {u + c : u e Ma,
c is any constant},
and denote by ua 6 Ma a minimizer of Ja satisfying / e"° = 4ir, then define (3.5)
F{a) = -!- /
v # • Vz e"° da.
47r y S 2
This way we obtain a multi-valued map F: B —> R3. Now the problem is to solve the equation (3.6)
F{a) = 0
a G B,
according to Theorem 2.1. For this purpose, we shall show that the map F has the Brouwer degree at zero point being the same as the following map (3.7)
G(x) = S7R{x) • sjx - AR(x)x,
xeS2,
119 provided that F is continuous in the interior of B (Lemma 5.2), then verify the continuity when the function R is close to 2 (Lemma 5.5). That means the equation (3.6) has a solution if deg(G, B, 0) ^ 0, and the solvability of (1.3) is obtained in this case. Moreover, based on some estimates for the solutions to (1.3) in [3], and some further calculus for degree, we remove the restriction that R is close to 2 and arrive at the following T h e o r e m 3.1 Suppose R satisfies | v R\ + I A R\ ? 0. If deg(G, B, 0) ^ 0, so the equation (1.3) has a solution, where the map G is denned by (3.7). Here we remark that, very obviously the map G does not vanish on 5 2 that is the boundary of B under the condition (3.1), so the Brouwer degree of G at 0 is well-defined. Below we change Theorem 3.1 into a simpler form Theorem 3.2. Let N = (0,0,1) be the north pole. Identify S2\{N} with the equatorial plane 2/1,2/2 via the stereographic projection: y : S2\{N} -* R2 xeS2,
(3.8)
1/1 = : ^ - ; V2 = j ^ - , 1-X3 l-x3 with its inverse transformation ,,. (3.9)
22/! Xi
=
• , , , , ,2
1 + I2/I
22/2 %2 =
,
,
I 122
1 + I2/I
|2/|2-1
X
3
-
•,
, |
12'
1+ bl2
and y = 00 corresponding to N. Then (3.10)
50 = dx\ + dx\ + dx\ = ———(dyl
+ dyl).
Now R and AR can be also viewed as functions of y in R2, and V ^ be viewed as a map from R2 to R2. Without loss of generality, we assume S7R(N) = 0,
120
which guarantees that as y —> oo,
Introduce G* : R2 ->• it 3 by (3.12)
G*(i/):=(vii(l/),-Ai2(y))
w
V 2
dyi'
2
Vy G i? 2 ,
9^'
^7
Because of (3.11), G*(y) ->• (0, 0, - A i?(7V))
as j/ -> oo,
where AR(N) ^ 0(by assumption (3.1) and V-R(N) = 0). We see that G* is a continuous map from S2 — R2 U {oo} into i? 3 \{0}. Thus ygm- continuously maps S2 into S2. Using Theorem 3.1, we can prove the following T h e o r e m 3.2. Suppose (3.1) holds. In the case AR(N) < 0, if deg(|^rr) ^ 1, then (1.3) has a solution; in the case AR(N) > 0, if" deg(j^rj) ^ —1, then (1.3) has a solution. Here G* is defined by (3.12). We shall give the main idea of proofs for Theorem 3.1 or Theorem 3.2 in section 5. In Theorem 3.1 or Theorem 3.2, the map G or G* has very simple expression: explicitly and simply depends on R, in fact, only on \/R and AR. It is not hard to calculus their degree. In the following, we give a few corollaries. Corollary 3 . 1 . Suppose (3.1) holds. If AR(x) A R(-x)
- yR{x)
• \/R(-x)
> 0 Vx £ S2,
then the equation (1.3) has a solution. Proof. Set Rt(x) = R(x) + tR(-x)
Vz e S2, te [0,1].
121
The corresponding map G = Gt {t e [0,1]), defined by (3.7) with R = Rt, satisfies |G ( (x)| 2
= | V R{x) - t v R{~x)\2 + (AR{x) + t A R{-x))2 > | V R(x)\2 + ( A E ( i ) ) 2 + 2t(AR(x) A R(-x) - \7R{x) • >\yR(x)\2 + (AR(x))2 >0
\/R(-x)),
by (3.1) and the condition AR(x) A R(-x)
- \/R{x) • yR(-x)
> 0.
So Gt has the degree independent of t by the homotopy invariant property. Observe that Ri(x) = R(x) + R(—x) is an even function and the corresponding map G\ is odd, hence has an odd degree by Borsuk theorem. We see that the map G = Go, corresponding to R, has the odd degree. Application of Theorem 3.1 immediately yields the conclusion. The proof is finished. When R is even, i.e. Moser's case, obviously AR(x) A R(-x)
- \?R{x) • \7R(-x)
= \ y R(x)\2 +
(AR(x))2,
being not negative, Va; G S2. Thus Corollary 3.1 is an extension for the even function cases. Corollary 3.2. Suppose (3.1) holds. If R only has non-degenerate critical points in {AR > 0}(or in {AR < 0}), and £ y e 5 + ( - l ) i n d ( y ) ± 1 ( or S y e S _ ( - l ) i n d W jt 1), then the equation (1.3) has a solution. Here 5+ := {yeS2:
VR(y)
= 0, AR(y) > 0}
S- := {yeS2:
yR(y)
= 0, AR(y) < 0}.
Corollary 3.2 extends Chang-Yang Theorem in which all critical points of R are assumed to be non-degenerate, so R is only permitted to have
122
isolated, finite number of critical points. In Corollary 3.2, R may have infinitely many critical points in {AR < 0}(or in {AR > 0}). Proof of Corollary 3.2. Denote by (u\, 112, U3) S R3, u\ + u\ + u\ = 1, the image of the map j^nrr. Take {111,112) as the coordinates at the north pole (0, 0, 1) and the south pole (0, 0, -1). By definition (3.12) of G*, \G*(y)\(Ul,u2)(y)
=vR{y) _ (i+\v\2 dR i+|y[2 dR\ ~ \ 2 dVl > 2 dy2 J •
Then it is easily seen that, at the point y satisfying S?R{y) = 0 and y ^ 00, dui
(3.13)
sign
du-2 dyi
sign
R.vwi ;
Ry R V2V1
Ry
= sign\D2 R(y)\
where D2R is the Hessian of R. Second! observe that for the map ^77, the inverse of the south pole Secondly, (0, 0, -1) is S+ = {y € S2 : vR(y) = 0, AR(y) > 0}, while the inverse of the north pole (0,0,1) is
S_ = { i / e S 2 : v % ) = 0,A%)<0}. If we suppose in S+ the function R is non-degenerate, the determinant in (3.13) is not vanishing with the sign equaling sign (-l) 11101 ^) for y ^ 00, y £ S+. Choose a maximum of R as N, i.e. R(N) = maxs2 R, then Ai?(JV) < 0 and S+ does not contain 00. It follows that the determinant in (3.13) is not vanishing for y 6 S+, then the south pole (0,0, —1) is regular. Moreover, (3.14)
G* d e g ( — - ) = E y e s + sign I" I
du\
8m
01X2
dyi 9U2 9j/2
= SyeS+(-l)ind^).
On the contrary if we suppose R in S- is non-degenerate, we choose a minimum as N, i.e. R(N) = mins2 R, then AR(N) > 0, implying 5_ does not contain 00. Similarly as above we see that the north pole (0,0,1) is regular. Notice that, if (ei,e 2 ) is an orienting basis for R2, an orienting
123
basis for S2 at the south pole (0,0, - 1 ) is (ei,e2), while at the north pole (0,0,1) (ei,— e 2 ) is orienting. So (3.15)
d e g ( - ^ r ) = - S y e s _ sign
dui 9yi du2 9yi
du\ dy2 du2 9y2
_-,und(y) = -Ss6S_(-l)
With (3.14) and (3.15), we apply Theorem 3.1 and immediately obtain the conclusion. The proof is finished. Let us go to the last corollary. Denote (3.16)
0 := {y € S2 : AR(y) > 0}.
Denote by g* the restriction of gradient of R on the boundary <9ft, i.e.,
Since AR = 0 on dfl, by (3.1) g* is a map into -R 2 \{0}, thus ^ where S 1 is the equator S1 = {u € R3 : |u| = 1, U3 = 0}.
: dSl -> S1,
Corollary 3.3. Suppose (3.1) holds. If A T extends to a map: Cl -» S1 (or to a map: S2\Cl -> S 1 ), then the equation (1.3) has a solution. Proof. Let /i : f) —> S1 be a continuous extension of Arr. Since both I^TT (with G* defined by (3.12)) and h map ft into the lower hemisphere {«3 < 0} with the same boundary value T^T, and since the hemisphere can be viewed as R2, there is homotopy, i.e. a continuous H : [0,1] x fi —> S2, satisfying ff(0,-) = | | ^ ; and
H(1, •) = /»,
9* G* H{t,-)\ea = | ^ - | = | ^ j |an-
By defining
in[0,l]x52\n,
H=~
we get a continuous H : [0,1] x S2 —» S2, which is an extension of H, satisfying C* U H(0,-)=
iG*r
124
so T^prr is homotopic to H(l, •) : S2 —> S2 which is an extension of h. Notice that T^rr maps S2\Cl into the upper hemisphere, then F(l,S2)c{u3>0} and hence H(l, •) has the degree 0. Thus deg(j^7j) = 0. In another case when A T extends to a map h : S2\Cl —> S1, a similar argument as above shows that Mpr\ is homotopic to a continuous h : S2 —>• S2 with h{S2) C {u3 < 0}, implying deg(jg^j) = 0 too. By Theorem 3.2 we finish the proof. The assumption in Corollary 3.3 that -PTT extends to Cl (or to S2\Cl) is equivalent to deg(j^|9n') = 0 10 I where Cl' is any connected components in Cl (or in S2\Cl), according to a fundamental theorem in topology, in the case where A is a manifold. For example, if R is an axi-symmetric function, for any annulus type connected components Cl' in Cl (or in S2\Cl), the degree of Arr\dQ' is vanishing. To the best of my knowledge, there seems no any counterpart of Corollary 3.3 in the past.
§4. Preliminaries In this section, we list some basic inequalities and estimates about the functional J and solutions of equation (1.3). First introduce some notations. For every a £ B, define a function
Mx)
= l g
° (CoS2l/x2Sin2l)2
XGS2
where 6 = d(x, r M the distance on (5 2 , q0) between x and -A-, and A = v a |a| ' \a\ Aa £ (0,1] is uniquely determined by the equation P((f>a) = a.
125
Set (4.2)
I(u) = l [ \yu\2dA 2 Js2
u^H^S2).
+ 2 [ udA, Js2
For ( 6 S 2 , ( 5 > 0 , denote (4.3)
Cc^ueff1^2);
Q(u) = (,d(u) = 8}
where (4.4)
Q(«) = P(«)/|P(u)|,
d(u) = | Q ( t t ) - P ( « ) |
if
P(u)^0.
Lemma 4.1 ([15], [23]). (1). For a € B there exists a unique conformal transformation Ta : 2 S -J- S 2 such that dy = e^adx
if
y = Ttt(ar).
(2). For a E B, <j)a has the properties: / e^° = 47r;
and
<^>a € Ma.
(3). We have /" eu < 4TT exp J(u)
tf1^2),
Vu €
with the equality if and only if u = <pa for a £ B. Lemma 4.2 ([8]). Let {ui}, i = 1,2, • • •, be a sequence satisfying | P ( U J ) | < 5 < 1 and J(tii) < c < +oo, then there exists a constant C, depending only on 1 — S and c, such that f \yUi\2dA
Vi = 1,2,3- ••.
Lemma 4.3. (Corollary 5.1 of [5]). Suppose u £ C^ts with i"(u) = 0(<5/3) for some y3 > 0 and 5 sufficiently small, / 2 eudA = An. Then for every function h e
C2(S2),
Js
^ J2 /ieudA = MO + ^ Aft(C)5+ 0(5/{- logS)).
126
Lemma 4.4. (Theorem 2(a) of [3]). Suppose R is a smooth function on 5 2 satisfying (1.3) with 0 < m < R < M, then there exists a constant C\ > 0, depending on m,M and min{| A R(x)\ : \jR{x) = 0}, such that for all solutions u to (1.3), |u| < C\. For proving our Theorem 3.1, the above Lemma 4.4 can not be directly applied. We need the following Lemma 4.5, a modified estimate. Let X the Banach space X = {u£C2'a(S2):
fu = 0},
and Ls : X -» X defined by (4.5)
Lsu = A " 1 (2 -
8?r
Rs eu)
Vwel
u
Rse where (4.6)
V 0 < s < 1, x G 5 2 .
R,(x) = 2 - 2s + sR(x)
Lemma 4.5. Suppose (3.1) holds. Then the set S = {ueX
: u - Lsu = 0, 0 < s < 1}
is bounded in X. Proof. If u £ S, for some s £ (0,1] we have (4.7),
-Au + 2-
&n
u
Rseu
=0
onS2,
I Rse
then, for a constant cu, v := u + cu satisfies (4.8),
- A w + 2 - t f , ev = 0
Since / u — 0, we see c u = ^
onS2.
v, the bound of u implies the bound of u,
thus it suffices to show that the set of all solutions to (4.8), (0 < s < 1) is bounded in C2,a.
127
Above Lemma 4.4 says that all solutions u to (1.3) have the estimate: ||u||c2,c < C where the constant C depends only on mini?, maxi?, and min{| A R(p)\ : S7R{p) = 0}. If we directly apply Lemma 4.4, it can only yield ||w||C2,<» < Cs where Cs depends on s > 0. However, by carefully observing the proof of it in [3] and noticing that our function Rs has the special property: \/Rs = s y R, we can find that all solutions to (4.8)5 have an uniform C 2 ' a -estimate independent of 0 < s < 1. Indeed, on the contrary suppose that there are Sk > 0; Vk satisfying (4.8)Sfc, but H^feilc° ->• oo as A; -^ oo. The same as in [3](pp.209), we get a new sequence Wk satisfying ||t«fe||cri = o(l) (as k -> oo) and
| v ( f l . * ° l k ) • s/x ew" = 0
(4.9)
where tpk = Tah with a^ = (1 - \k)P for some fixed point P G S2 and Afc G [0,1); Afc -> 1. By ^RSk = SkSJ R, we see (4.9) becomes (4.10)
v{Ro1>k).s?xev"-=0.
/
It is exactly the formula (3.10) in [3] where it was shown that their (3.10) implies V-R(-P) = 0 and AR{P) = 0. So above (4.10) implies vR(P) = AR(P) = 0, contradicting to the assumption (3.1). The proof is finished.
§5. Main steps of proofs In this section we give the main steps of proving Theorem 3.1 and Theorem 3.2. We omit the details and proofs of all lemmas. Main idea for proof of Theorem 3.2. Using y as the stereographic projection (3.8) and (3.9) from in S2\{N} to the equatorial plane 2/1,2/2Writing
x=\
x2
x* J
;
G*(y)=\
M™
\ -AR
J
(x\,X2,Xz)
128
Denote A
yy>
-
\
2
9j» '
2
X
dy2 >
)
( 1 + 2/1 — 2/i -2j/i2/2 -22/12/2 1 + y\ ~ 2/2 V 2^ 2y2
= i+fcp
Then by definition of G and G*, it follows (5.1)
Vx 6 S2;
G(x)=A(y)G*(y)
y 6 R2.
Here the matrix A(y), y £ R2, is easily verified to be orthogonal with A2 =id. It should be mentioned that in (5.1), the matrix A is not continuous on S2 (discontinuous at the infinity, the north pole N), however it is easy to see that if a continuous / : S2 -¥ S2 maps the north pole N to N or to the south pole (0,0,-1), then Af : S2 -> S2 is a continuous map. That is a reason why we always assume \?R(N) — 0. Let k be an integer, and introduce Ik : S2 —>• 5 2 by
(
2r cos fc(/> 2r sin k<j> r2-l
where (r, <j>) being the polar coordinates: yi—r cos (j>, 2/2 = r sin (/>,
Vr > 0, 0 < <j> < 2ir.
Assume deg(T^rr) — m. We prove the following (1) in the case AR(N) < 0, deg(G, B, 0) = (2)in the case AR(N)
deg(AIm);
> 0,
deg(G, B, 0) = -deg(Afi),
with / = - m ;
(3)for any integer, it is true deg^J*) = 1 - k.
129
With these facts (l)-(3), we obtain deg(G, B, 0) = 1 - deg( | | 1 )
if A R(N) < 0;
deg(G, B, 0) = - 1 - d e g ( | | ^ )
if A R(N) > 0.
Then Theorem 3.2 is obtained by Theorem 3.1. Main idea for proof of Theorem 3.1. Step 1. Calculus the degree of the map F, under the assumption of the continuity. Here F is defined by (3.5), i.e. F(a) = — / V-R • V z e"° dx 4TT JS2
aeB
with ua G Ma being a minimizer of J a satisfying / 2 e"° = 4-7T. is In this step, a key point is the following Lemma 5.1. Let / € C3(S2). < x, \jf{x)
Then Vz G S 2 ,
• y z > = 0;
< x, A ( V / ( z ) • v z ) > = - 2
A
/(a?)-
Here < •, • > is the inner product of R3. Applying the asymptotic formula Lemma 4.3 we first show that the map F has the expression: when 5 > 0 sufficiently small and a G B with \a\ = 1 — 6, F(a)
=
4^r /
v i ?
'v x
= V-R(a;) • V
x
+\
e
"° A
^
(VR(X) • Va;)<5 + o((5).
Secondly, using Lemma 5.1 and the condition (3.1), we can construct a homotopy H : [0,1] x S2 such that H never vanishes and satisfies H(0,-)=F;
130
H(l,x)
= V-R- S/x-
xeS2.
ARx
Then it proves the following Lemma 5.2. Suppose (3.1) holds. There exists e > 0, depending on ||.R||C2, such that d e g ( F , B r , 0 ) = deg(G, B, 0)
when 1 - e < r < 1,
provided that the correspondence a H-> ua is a continuous map from Br to H1. Here G is defined by (3.7). Step 2. Establish an inequality of Poincare type in some subspaces of H1. It will play a fundamental role in later step 3 and step 4. We recall that, for a € B, ua € Ma is a minimizer of Ja satisfying /
Js2
eUa = A-K, it also depends on the function R, and TUaMa denote the
tangent space of Ma, at ua, i.e. the set of functions v € H1 satisfying / eUa(x - a)v = 0) (see (2.8)). Let Ta : 5 2 -»• S2 the unique conformal transformation determined by Lemma 4.1(1). We observe, from Lemma 4.1(3), that ua =
By using Lemma 5.3 and some careful calculus, we can prove Lemma 5.4. Let 0 < r$ < 1. There exist 5 > 0 and p, > 2 such that, when |i? — 2| c o < 5, there holds
/ l v < f > M / l«°77T
is 2
Js2
131
for every a G BTQ and v G TUaMa. Step 3. Show the continuity for the correspondence a 4 « 0 when R is close to the constant 2. It suffices to show the uniqueness of minimizers of Ja subject to / 2 e"° = Js An. As we mentioned in step 2, obviously it is true when R = 2. For R close to 2, we can prove the uniqueness based on Lemma 5.3 and Lemma 5.4, that is the following L e m m a 5.5. Given e G (0,1), there exists 6 > 0 such that, when \R — 2|c?o < S, for every a G Bi_ e , the functional Ja has minimum at unique ua G Ma with / eUa = 4ir. Combining Lemma 5.2 and Lemma 5.5, we have proved Theorem 3.1 in the case R is close to 2, by Theorem 2.1. Moreover, in order to remove the restriction, we need to pay attention to the Leray-Schauder degree of a map related to equation (1.3), still in the case R close to 2. It is what we will do in the next step. Step 4. Let X the Banach space X = {u £ C2'a(S2)
: / u = 0}, and Ls : X ->
X defined by (4.5) and (4.6), i.e. Lsu = A " 1 (2 -
8?r
/Rse
Rs eu)
V«£l
u
where Rs(x) = 2 - 2s + sR(x)
V0 <s < 1, x G S2.
We shall compute the Leray-Schauder degree deg(i — Ls, fi, 0) with s close to 0, where the set ft c I is taken to be an open bounded set that contains S, the set of zero points of (I — Ls), Vs G (0,1], and S D dfi. = 0 ( by Lemma 4.5). In fact, we wish to prove that for some sufficiently small si > 0, (5.2)
deg(7 - LS1, n, 0) = ± deg(G, B, 0).
132
For this purpose, we first prove L e m m a 5.6. Given M > 0, there exists 6 > 0 such that, when \R 2|co < S, any solutions u of (1.3) satisfying ||u||c2,<» < M, must be the unique minimizer of Ja subject to / Reu = 8ir, where a = P(u). Js2 ^From this lemma we can see that, when s close to 0 (implying Rs close to 2), any zero point u of the operator (I — Ls) must be the unique minimizer of Ja subject to / Rseu = 8ir where a = P(u). Based on this knowledge about the zero points, and based on Lemma 5.3 and Lemma 5.4, we can construct a series of homotopy H(t, •), t > 0, that never vanishes on the boundary Oil and deform the restriction (/ — Lai)\aa, for some sufficiently small si > 0, to a map like G x I: S2 x U, with U an open set of a subspace in X of codimension 3. This leads to the above formula (5.2). Step 5. Complete the proof of Theorem 3.1. The Leray-Schauder degree deg((7 — Ls), CI, 0) is well defined and independent of s € [s\, 1]). Then from last step we have deg((J - LJ,
n, 0) = ±deg(G, B, 0) # 0.
This means that the operator (I — Li) has a zero point, which produces a solution of the equation (1.3).
References [1] T. Aubin, Meilleures constantes dans le theoreme d'inclusion de Sobolev et un theoreme de Fredholm non lineaire pour la transformation conforme de la courbure scalaire. J. Funct. Anal. 32, (1979), 148-174. [2] J. Bourguignon and J. Ezin, Scalar curvature functions in a class of metrics and conformal transformations, Trans. A.M.S., 301(1987), 723736.
133
[3] A. Chang, M.Gursky and P. Yang, The scalar curvature equation on 2and 3-spheres, Calc. Var. PDE 1(1993), 205-229. [4] K.C. Chang and J. Liu, On Nirenberg's problem, International J. Math. Vol.4, No.l (1993), 35-58. [5] A. Chang and P. Yang, Prescribing Gaussian curvature on S2, Acta Math. 159(1987), 215-259. [6] A. Chang and P. Yang, Conformal deformation of metric on S 2 , J. Diff. Geom. 23(1988), 259-296. [7] A. Chang and P. Yang, A perturbation result in prescribing scalar curvature on Sn, Duke Math. J., Vol.64, No.l(1991), 27-69. [8] W. Chen and W. Ding, A problem concerning the scalar curvature on S2, Kexue Tongbao 33(1988), 533-537. [9] W. Chen and W. Ding, Scalar curvatures on S2, Trans. A.M.S., 303(1987), 365-382. [10] W.Chen and C. Li, A necessary and sufficient condition for the Nirenberg Problem, Comm. Pure and Appl. Math. XLVIII(1995), 657-667. [11] W. Chen and C. Li, A note on Kazdan-Warner type conditions, J. Diff. Geom. 41(1995), 259-268. [12] K.S. Cheng and J. Smoller, On conformal metric with prescribed Gaussian curvature on S 2 , Trans. A.M.S., 336(1993), 219-255. [13] J. Escobar, and R. Schoen, Conformal metrics with prescribed scalar curvature, Invent. Math. 86(1986), 243-254. [14] Z.C. Han, Prescribing Gaussian curvature on S2, Duke Math. J. 61(1990), 679-703. [15] C.W. Hong, A best constant and the Gaussian curvature, Proc. A.M.S., 97(1986), 737-747. [16] C.W. Hong, A note on prescribed Gaussian curvature on S2, J. Partial Diff. Equa.,Vol.l. No.l(1988), 13-20. [17] M. Ji, On symmetric scalar curvature on S2, Chin. Ann. math. 20B(1999), 325-330. [18] J. Kazdan and F. Warner, Scalar curvature and conformal deformation of Riemannian structure, J. Diff. Geom. 10(1975), 113-134. [19] J. Kazdan and F. Warner, Curvature functions for compact two manifolds, Ann. Math. 99(1)(1974), 14-47. [20] J. Kazdan and F. Warner, Existence and conformal deformation of metric with prescribing Gaussian and scalar curvature, Ann. Math.
134
[21] [22] [23] [24] [25]
101(2)(1975), 317-331. J. Moser, A sharp form of an inequality by Neil Trudinger, Indiana Univ. Math. J. 20(1971), 1077-1092. J.Moser, On a nonlinear problem in differential geometry, In: Dynamical Systems, Academic Press, New York (1973), 273-280. E. Onofri, On the positivity of the effective action in a theory of random surface, Comm. Math. Phys., 86(1982), 321-326. R. Schoen and S.T. Yau, Differential Geometry, Monograph in pure and applied mathematics, No.l8(1988), Science Press. X. Xu and P. Yang, Remarks on prescribing Gauss curvature, Trans. A.M.S., 336(2)(1993), 831-840.
P E R I O D I C M E A N CURVATURE A N D BEZIER CURVES
K. KENMOTSU Mathematical Institute, Tohoku University 980-8578 Sendai, Japan E-mail: kenmotsu@math. tohoku. ac.jp We show a recipe to get the periodic surface of revolution in the Euclidean three space such that its mean curvature is a given function.
1. Introduction It is well known that the profile curves of surfaces of revolution with constant mean curvature in R3 are constructed by the rollings of ellipses or hyperbola along the axis of rotation (Delaunay x ). In the recent paper by Kenmotsu 3 , he showed a new method to have such profile curves: In fact a circle with radius r produces one parameter family of surfaces of revolution with constant mean curvature l/2r. Moreover, generalizing this, he found all profile curves that make periodic surfaces of revolution and proved that these are provided by a class of Bezier curves. The purpose of this paper is to show the figures of these periodic profile curves and the commands of Mathematica which is a symbolic manipulation program by Wolfram. Moreover, the resemblance and the difference between those periodic surfaces and Delaunay surfaces are clarified. This paper3, extends the manuscript of my talk in the International symposium on differential geometry held in Shanghai on September 2001. 2. Periodic surfaces of revolution Let C = (x(s),y(s)) be a complete smooth curve parametrized by arc length, which generates the surface of revolution S in R3 as follows : S= {(x(s),y(s)
cos6,y(s) sin0) G R3 | s G R, 0 < 6 < 2TT} ,
where x-coordinate of R3 is the axis of rotation. C is called the profile curve of S. "Dedicated to Prof. Su Buchin on his centennial birthday
135
136
The surface of revolution S is called periodic if the coordinate function y(s) is periodic, that is, there is a positive number L such that y(s + L) — y(s) for all s E R. L is said to be the period of S. The mean curvature function of a periodic surface of revolution is periodic, but the converse does not hold in general, catenoid giving us the counter example. We have a necessary and sufficient condition for a continuous periodic function to be the mean curvature of some periodic surface of revolution by Kenmotsu 3 . To state this, given a continuous function H = H{s) on R, we put h(u) = = 22 / " H(s)ds, ueR. (1) Jo Jo Theorem 1. Let H(s) be a continuous periodic function on R with period L. Then, the function H(s) is the mean curvature of a periodic surface of revolution S with period L if and only if it satisfies the condition : Jo cos(h(u))du sin h(L)
=
JQ£ sin(h(u))du 1 - cos h(L) '
[
'
Convention. A denominator in Eq. (2) vanishes if and only if the corresponding numerator also does. We show now a recipe to get periodic profile curves : Let H(s) be a periodic function on R satisfying the criterion Eq. (2). Put F(s) = / sin(h(u))du, Jo
(3)
G{s) = [ cos(h{u))du .
(4)
Let c be the absolute value of the common ratio of Eq. (2) when some of the terms do not vanish and any positive number otherwise. Under these notations, we have a periodic profile curve C = (x(s;c),y(s;c)) due to Kenmotsu 3 as follows : y(s;c) = {(F(s)-c)2+G(s)2}i, V
;
Jo
(5)
{ ( F ( s ) - c ) 2 + G( s )2}l
V
'
This curve C is parametrized by arc length and passes through (0, c). When a periodic function H(s) trivially satisfies Eq. (2), we have a one parameter family of these curves with the same mean curvature which is given by {(x(s;c),y(s;c))eR2
| c > 0}.
137
When a periodic function H(s) non-trivially satisfies Eq. (2), for any c'(^ c) the curve (x(s;c'),y(s;c!)) generates a non-periodic surface of revolution with periodic mean curvature H(s). Hence, in this case, we have only one profile curve which is periodic.
Example 1. (Torus) Put ,. . /(S)
1/ , coss 1+ = 2l" 2^Ts)
\ •
-°°<«<°°-
(7)
It non-trivially satisfies Eq. (2) with L = 2TT and c = 1. In fact, this is the mean curvature of a torus in R3 whose profile curve is a circle and passes through (0,1). Since this function f(s) non-trivially satisfies the criterion Eq. (2), there is just one value c, in this case c = 1, such that the profile curve (x(s; l),y(s; 1)) through (0,1) is periodic, but for any other c (7^ 1) the profile curve (x(s;c),y(s;c)) is not. Figure 1 shows the curve {G(s), F(s)) made from f(s) and Figure 2 the profile curves (x(s;c),y(s;c)), -87r < s < 8n made from these G(s) and F(s) when c = 0.5, 1, 2.
Figure 1. The curve (G(s),F(s)) of a torus.
-2-1
1 2
-1-0.5 '
0.5
1
-4
-2
Figure 2. Profile curves with the same mean curvature We remark that the direct use of the command of Mathematica for the integration of Eq. (7) may not work well, because it is an improper integral. We integrate the Fourier expansion of Eq. (7) to prove Eq. (2). Let us consider the plane curve (G(s),F(s)), which is parametrized by arc length and has curvature 2H(s). This means that from the profile curve of a surface of revolution with mean curvature H(s), we have a plane curve
138
with curvature 2H(s). Reckziegel was first to remark that conversely, from a plane curve T with curvature k(s), a profile curve C whose surface of revolution has mean curvature k{s)/2 is constructed (See the first paper by Kenmotsu 2 .) Definition 1. (G(s),F(s)) is called the source curve of a surface of revolution S with mean curvature H(s). The curve (x(s;c),y(s;c)) is called to be associated with the source curve (G(s),F(s)). In section 3 we explain how to get periodic profile curves made from curvatures of closed plane curves and in section 4 we do how to make periodic profile curves from curvatures of some non-closed curves, in which Bezier curves are used. 3. Closed curves Let T be a twice continuously differentiable closed plane curve parametrized by arc length and L the length. The curvature of T, k(s), is considered as a periodic function on R with period L. We may assume that Y starts from the origin of (x, i/)-plane and is tangent to x-coordinte at the origin. It follows from the fundamental theorem of smooth curves theory that, on s € [0,L], we can write it as r = ( / cos ( /
k(t)dt\ du, J sin ( /
k(t)dt j du) .
(8)
With the famous theorem for the total curvature of closed plane curves, this formula implies that the function k(s)/2 trivially satisfies Eq. (2) with period L, because all terms in Eq. (2) vanish. Therefore, for any positive number c, there exists a periodic profile curve C(s; c) such that the resulting surface of revolution is periodic with period L, has the mean curvature k(s)/2 and passes through (0,c). We have seen that, a smooth closed curve with curvature k(s) provides the 1 parameter family of periodic profile curves {C(s;c) | 0 < c < oo}, where k(s)/2 is the mean curvature of resulting surfaces of revolution. This resembles Delaunay surfaces. The deformations of such periodic profile curves are now shown by examples : Example 2. (Circle) Let T be a circle with radius r centered at (0, r) e R2. The resulting {C(s;c) | 0 < c < oo} is the set of profile curves of De-
139
launay surfaces such that l/2r is the mean curvature of the corresponding surfaces of revolution. See Figure 3.
350 0.25
S.8D571.15 6 Figure 3. The deformation of profile curves made from circle We remark that this method is different from Delaunay 1.
E x a m p l e 3. (Ellipse) Let T be an ellipse such that one of the vertex is tangent to x-coordinate at the origin of (x, i/)-plane. The resulting 1parameter family of periodic profile curves resembles profile curves of Delaunay surfaces. See Figures 4 and 5.
5
2
10
4
15
6
Figure 4. The deformation of profile curves made from ellipse
140
Figure 5. The profile curves of surfaces of revolution associated with an ellipse. E x a m p l e 4. (Limagon) Let F be a Limagon given by (2acos[t] + 6)(cos[t],sin[t]),
— oo < t < oo,
where a and b are real constants. Figure 6 shows a Limagon, its curvature, the corresponding profile curve, and the resulting surface of revolution.
Figure 6. Lima^uii and the- associated bin face of n-volution Now we take a rose curve as F. Here is the command of Mathematica to draw the pictures of periodic surfaces of revolution associated with the rose curve. The rose curve is given by the command of a [ t _ , n _ , b _ ] := {2 * b * Cos[t] * Cos[n* t],2 * b * Cos[t] * Sim[n* t]};
141
Putting n := 2 ; b := 1/3 ;, Figure 7 shows the rose curve, which is given by the command of ParametricPlot[Evaluate[a[t, n, b]], {t, 0 , 2 * Pi}, AspectRatio —> Automatic]
Figure 7. The curve a[t, 2,1/3] Next we have the command to compute the curvature curve.
K.1[£]
of the rose
s'[t_] :=Simplify[(D[a[t,n,b],t].D[a[t,n,b],t]) 1 / a ] ; J[{pl_,p2_}]:={-p2,pl}; /cl[t_] :=D[a[tt,n,b],{tt,2}].J[D[a[tt,n,b],tt]]/s'[t] 3 /. tt -»• t ; nl[t] above is the curvature of the rose curve for the parameter t which is not arc length. Let s = s(t) be the transformation of parameters from t to arc length s. To get the function F(s) of Eq. (3), where s is arc length, we use the differential equation :
£ = - . ( / « i ( . ( . ) ) ! * ) ! . "(»)-»• We put k[t_] := Simplify[«l[tt] * s'[tt]]/.tt -> t ; Since this function is rather complicated to treat by Mathematica, we use the command of interpolation as follows : datal = Tableau, k[u]},{u, 0 , 2 * Pi + 0.1, 0.02}]//N ; K2 = Interpolation[datal] ; Now, we get F(s) = ff[s] and G(s) = gg[s] by /•tt
soil = NDSolve[{f [tt] = = Sin
/ 7o
«2[u]du *s'[tt],
142
/
g'[tt] = = Cos
re2[u]du *s'[tt], f[0] = = 0, g[0] = = 0},
{g,£},{tt,G,2*Pi}][[l]] ; ff[s_] := f[s]/. s o i l ; gg[s_] := g[s]/. s o i l ; Putting c = 1.5, we define y-coordinate of the profile curve through (0, c) by y[8_,c_]:=((ff[s]-c)a + gg[8]3)1/3; The next command computes ^-coordinate of the profile curve through (0,c). sol2 = NDSolve[{X'[t] — = I gg[*] * Sin
/
«;2[u]du
(ff[t] - c) * Cos
X[0] = = 0},{X},{t,0,2*Pi}][[l]] ;
/ K2[u]du J /y[t,c],
x[t_,c_] := X[t]/. so!2 ;
7 [ t - , c _ ] = {x[t,c],y[t,c]} ; Now, we have the figure of a profile curve through (0,1.5) which is made from the rose curve by : p i = ParametricPlot[7[t, c]//Evaluate, {t, 0,2 * Pi}, AspectRatio —»• Automatic, DisplayFunction —>• Identity] p2 = ParametricPlotpyjt, c] + {*y[2 * Pi, c][[l]], 0 } / / Evaluate, {t, 0,2 * Pi}, AspectRatio —J- Automatic, DisplayFunction -4- Identity] ; Show[pl,p2, DisplayFunction -*• §Display Function] Figure 8 shows the profile curve y[t, c], 0 < t < Ait, and the corresponding surface of revolution. By changing the constants n, b and c used here we find other profile curves.
8
">4/
Figure 8. The profile curve and the resulting surface of revolution associated with a Rose curve.
143
4. Bezier curves In this section we treat the nontrivial case of Eq. (2) such that there are non-zero terms in Eq. (2) . The Eq. (2) says that the coordinates of the end point of the curve (G(s),F(s)) are related to the tangent of the point by
2B. =
F
^
(9)
V F'(L) 1 - G'(L) ' We put po = (0,0), and let pi = (rcos8,rsin6),r ^ 0, be any point of R2 - {(0,0)}. Let us consider a C2 curve segment r 0 = (G(s),F(s)), 0 < s < L, parametrized by arc length which starts from po and terminates at Px such that the both tangent vectors at the end points are (G'(0), -^'(0)) = (1,0), and (G'(L),F'(L)) = (cos26,sin26). Moreover, we assume that the curvatures of the curve To at the both end points have the same values. We extend To by some isometries of R2 so as the resultant curve T is smooth, i.e., class C 2 , and complete. The curvature of T, k(s), is a continuous periodic function on R with period L and the function k(s)/2 satisfies Eq. (2) with period L. The source curves made above may not be closed and the periodic surfaces associated with the curves are isolated. That is, we have the special value c such that the resultant periodic curve starting at (0, c) is periodic, where the constant c is given by the common ratio in Eq. (2) and for any other c'(^ c), the profile curve starting at (0,c') is not periodic. Bezier curves are now used to find these source curves T0 from the given data of ends points. Let B(i), 0 < t < 1, be a Bezier curve with the control points bo,bi,...,b„ such that bo = (0,0),bi = (b, 0), ...,b n _i = (-6cos26' + cot 6, -bsin20 + 1), b„ - (cot 0,1), TT/4 < 6
B e r n s t e i n [i_Integer, n _ I n t e g e r , t_Integer] := Binomial [n, i] * t1 * (1 - t ) n _ i ; Bernstein [i_Integer, n_Integer]
144 = M a p [ B e r n s t e i n [ # , n , t ] & , R a n g e [0,n]] We define the Bezier curve with control points p t s by B e z i e r [ p t s _ , t _ ] := B e r n s t e i n [Length[pts] — l , t ] . p t s ; Let us take t h e following points, as a n example, p t s l = { { 0 , 0 } , { 0 . 4 , 0 } , {0., 1 } , { 1 , 0 . 6 } , { 1 , 1 } } ; By the command of B e z i e r [ p t s l , t ] , we have a parametric form of t h e Bezier curve with the 5 control points. T h e left hand side of figure 9 is obtained by N e e d s ["Utilities 'Filter-Options'"]; Options [BezierCurve] = {ShowPoints —¥ False}; BezierCurve [pts_,opts
] := Module [{t,ptsQ = S h o w P o i n t s / . { o p t s } / .
Options[BezierCurve]}, Show[Block[{§DisplayFunction = Identity}, ParametricPlot[Evaluate[Bezier[pts, t]], {t, 0 , 1 } , Evaluate[FilterOptions [Par ametricPlot, opts]], P l o t S t y l e ->• Thickness[0.001]]], Graphics[If [ptsQ, PointSize[0.02], Map[Point,pts]},{}]], FilterOptions[Graphics, opts],PlotRange -> AH]] ; b l = BezierCurve[ptsl, ShowPoints —> True, Frame —> True, A x e s —> False, P l o t R a n g e —>• {0,1.0}, AspectRatio —>• Automatic] In order to have the right hand side of Figure 9, we compute the curvature of the Bezier curve.
Figure 9. Bezier curve with 5 control points and the surface of revolution associated with the Bezier curve.
s l [ t _ ] :=Simplify[(D[Bezier[ptsl,tt],tt].D[Bezier[ptsl,tt],tt]) 1 / 2 ]/.
145 t t -»• t; J [ { p l _ , p 2 _ } ] := { - P 2 , p l } ; ic[t_] := D [ B e z i e r [ p t s l , t t ] , { t t , 2 } ] . J [ D [ B e z i e r [ p t s l , t t ] , t t ] ] / s l [ t ] 3 / . t t -» t ; d a t a 2 = T a b l e a u , /c[u] * sl[u]}, {u, 0 , 1 , 0.02}]//N ; K2 — I n t e r p o l a t i o n [ d a t a 2 ] ; We get F(s) = ff[s]
and G(s) = gg[s] by
s o i l = NDSolve[{f'[tt] = = Sin[Integrate[K2[u], {u, 0, tt}] * s'[tt], g'[tt]
= = Cos[Integrate[K2[u],{u,0,tt}] *s'[tt],f[0] = = 0,g[0] = = 0},
{g, f } , {tt, 0,1}])[[1]] ; ff[s_] := f [s]/. s o i l ;
gg[s_] := g[s]/. s o i l ;
Let us define y-coordinate of the profile curve through (0,1) by y[s_] := ((ff[s] - l ) 2 + ( g g H ) 2 ) 1 / 2 ; sol2 = NDSolve[{X'[tt] = = (gg[tt] * Sin[Integrate[/c2[u], {u,0, tt}] ~(ff[tt] - 1) * Cos[Integrate[/c2[u], {u, 0, tt}])/y[tt], X[0] = = 0}, {X}, {tt,0,l}])[[l]]/. t t - M ; ; we define x-coordinate of the profile curve by x[t_] := X[t]/. sol2 ;
c[t_] = {x[t], y[t]} ;
The profile curve is depicted by the commands of p i = P a r a m e t r i c P l o t [ c [ t ] / / E v a l u a t e , {t, 0 , 1 } , A s p e c t R a t i o —>• A u t o m a t i c ] p 2 = P a r a m e t r i c P l o t [ c [ t ] + {c[l][[l]], 0 } / / E v a l u a t e , {t, 0 , 1 } , A s p e c t R a t i o —• A u t o m a t i c ] ; S h o w [ p l , p 2 ] Finally, we have the commands to get the shape of the right hand side in Figure 9 as follows: p 3 = P a r a m e t r i c P l o t 3 D [ { x [ t ] , y [ t ] * C o s [ u ] , y [ t ] * Sin[u]}, { t , 0 , 1 } , {u, 0, 2 * Pi}], P l o t P o i n t s —¥ 50, A s p e c t R a t i o —> A u t o m a t i c ] ; p 4 = P a r a m e t r i c P l o t 3 D [ { x [ t ] + c[l][[l]], y[t] * Cos[u], y[t] * Sin[u]}, {t, 0 , 1 } , {u, 0,2 * P i } , P l o t P o i n t s -»• 50] Show[p3, p4, B o x e d —¥ False, A x e s —> False, A s p e c t R a t i o —¥ A u t o m a t i c ]
Acknowledgments This reserch was supported in part by the Grant-in-Aid for Scientific Reserch (No.12440012), Japan Society for the Promotion of Science 2001.
146
References 1. C. Delaunay, Sur la surface de revolution dont la courbure moyenne est constante, J. Math. Pures. Appl. Ser. 1, 6(1841), 309-320. 2. K. Kenmotsu, Surfaces of revolution with prescribed mean curvature, Tohoku Math. J. 32(1980), 147-153. 3. K. Kenmotsu, Surfaces of revolution with periodic mean curvature, preprint, 2001.
Almost Complex Manifolds and a Differential Geometric Criterion for Hyperbolicity Shoshichi Kobayashi Department of Mathematics, University of California, Berkeley
0.
Introduction
The concept of almost complex structure was introduced by Ehresmann in 1947 [3], and the 1950s saw much activities in geometry of almost complex manifolds, including a major achievement by Newlander-Nirenberg [10] on integrability problem. The recent work of Gromov [5] and his school on symplectic geometry revived interest in almost complex structures; for every symplectic manifold admits a compatible almost complex structure, which can be profitably used to study the symplectic structure. The main difference between complex manifolds and almost complex manifolds is that a generic almost complex manifold does not admit, even locally, any nonconstant holomorphic functions. This difficulty coming from paucity of local holomorphic functions may be overcome by differential geometry, partial differential equations and hyperbolic complex analysis. We shall show here effectiveness of differential geometric methods in dealing with hyperbolicity. 1.
Almost complex manifolds
We recall that an almost complex structure on a real manifold M is an endomorphism J of the tangent bundle TM such that J o J = —I. Since the determinant det(—/) = (det J ) 2 must be positive, it follows that the dimension of M must be even if it is to admit an almost complex structure. We extend J complex linearly to the complexified tangent bundle TCM. The eigenvalues of J are ±i, and TCM decomposes (1.1)
TCM
= T'M + T"M,
where T'M and T"M correspond to the eigen values i and —i, respectively. We often identify TM with T'M by the map £eTM
->Z-iJ£e
147
T'M.
148
Under this identification, we have a natural orientation on M. If M is a complex manifold, then the decomposition (1.1) is naturally given, and setting J to be mutliplication by i on T'M and — i on T"M defines an almost complex structure on M. There are certain topological obstructions to the existence of almost complex structure. In general, the fc-th Pontrjagin class pk(M) of (the tangent bundle of) M is defined to be pk(M) = (-l)kc2k(TcM). If M admits an almost complex structure J, then c{TcM)
=
c{T'M)c(T"M)
=
(1 + d ( M ) + • • • + c n (M))(l -
C l (M)
+ • • • + (-l)"c„(M)).
It follows that (-l)kPk(M)
= (-l) 2fc c 2fc (M) + (-l) 2A =- 1 c 1 (M)c 2A _ 1 (M) + • • • +
In particular, if a sphere 5 structure, then 0 = Pm{Sim)
4m
c2k(M).
of dimension 4m admits an almost complex = 2c2m{Sim)
= 2e(5 4 m ) = 4,
which is a contradiction. This shows that a sphere of dimension 4m admits no almost complex structures. Actually, S2 and S 6 are the only spheres that admit almost complex structures; the proof involves more sophisticated topological arguments (see Kobayashi-Nomizu [9] for references). An almost complex structure on S 6 can be constructed by means of Cayley numbers, (S 6 is identified with the space of unit purely imaginary Cayley numbers), and an almost complex structure on S6 induces an almost complex structure on every hypersurface in the 7-dimensional Euclidean space via the Gauss map, see for example [9]. The compact homogeneous almost complex manifolds of positive Euler number have been classified (Hermann [6]). S 6 is the simplest such example. For 4-manifolds the existence problem has been completely solved. Namely, an oriented compact 4-manifold M admits an almost complex structure if and only if there is an element c £ H2(M; Z) such that (1.2)
c2 = 2e(M) + 3a(M),
c = w2(M)
(mod 2),
and c is given by c = Ci(M). Moreover, c = Ci(M) determines the almost complex structure uniquely up to a homotopy. For example, a product M of a compact orientable 3-manifold N with a circle S1 has e(M) = <J(M) = 0, and there is an almost complex structure such that c\{M) = 0; in this case, it is, of course, easy to construct an almost complex structure since the tangent bundle of N splits into a 2-plane bundle and a line bundle.
149
Other examples of manifolds which admit almost complex structures include even-dimensional parallelisable manifolds, symplectic manifolds, products of two contact manifolds. The existence of almost complex structures is a topological question; it is a problem of reducing the structure group GL(2n; R) of the frame bundle of a 2n-manifold to GL(n;C). However, the existence of complex structures is a much deeper question. It is a difficult problem to decide whehter a given (non-complex) almost complex manifold admits a complex structure or not. It is still an open question whether S 6 admits a complex structure or not. We recall that given two almost complex manifolds (M, J) and (M,J), a map / : M —> M is said to be holomorphic if it commutes with J and J, i.e., (1.3)
Jo/. =/.oJ.
In fact, this is the Cauchy-Riemann equations when J and J are complex structures. A holomorphic function is a holomorphic mapping into C. As I said earlier, for a generic almost complex structure on M, even locally there are no nonconstant holomorphic functions. Thus there is no complex function theory on almost complex manifolds. However, there is an abundant supply of holomorphic mappings from a disk of C into any almost complex manifold M. Let D be a unit disk in C. A holomorphic map / : D —> M is called a holomorphic disk. An important theorem of Nijenhuis-Woolf [10] says that given a point p G M and a tangent vector u G T'pM there is a holomorphic disk f:D->M tangent to u. Once we have the concept of holomorphic disk and the theorem of Nijenhuis-Woolf, we can immediately define the intrinsic pseduo-distance d,M and hyperbolicity for an almost complex manifold M exactly in the same way as in the complex manifold case, [7]. In Section 5, we discuss the hyperbolicity question using differential geometric means.
2
Canonical connections for almost Hermitian manifolds
An almost Hermitian manifold is an almost complex manifold (M, J) with a Hermitian metric g. By definition, g is a a Riemannian metric satisfying the condition (2.1)
g(Ju, Jv) = g(u, v),
u,v € TXM,
x G M.
150
Given an almost Hermitian manifold (M, J,g), it is natural to consider, among all affine connections on M, those which makes both J and g parallel, i.e., (2.2)
Vg -0,
V J = 0.
There are continuously many connections V satisfying condition (2.2). In order to make V unique, additional conditions are necessary. Let e\, • • •, e„ € T'M be a local unitary frame field. Let 81, • • •, 9n be the dual local unitary co-frame field; they are locally defined (l,0)-forms and g
= e1^ + ••• +
enen.
Then the connection form for a connection satisfying (2.2) is a locally defined 1-form w = (wj) with values in the Lie algebra u(n) of U(n); it is skew-Hermitian:
(2.3)
wj.+w?=0,
which is equivalent to (2.2). The first structure equation of the connection is given by (2.4)
d0i =
-^2u)ABj+ei,
and the second structure equation by (2.5)
d w ^ - ^ w j A w f + nj,
where © = (0 l ) and fi = (flj) are called the torsion form and curvature form, respectively. Since 0 and fi are 2-forms, they decompose into components of degree (2,0), (1,1) and (0,2).
(2.6)
0 = 0(2-°> + e*1-1* + 0 (o>2) , n = n (2 - 0) + n*1-1) + ft(0-2).
Among the connections satisfying (2.2), the most natural is the one defined by the following theorem of Ehresmann-Libermann [4] (for the proof, see [8]). (2.7) Theorem. Every almost Hermitian manifold (M, J, g) admits a unique affine connection such that J and g are parallel and the torsion 0 has no (1,1)-component. We call the unique connection defined by (2.7) the canonical connection of the almost Hermitian manifold (M,J,g). This connection introduced by Ehresmann and Libermann generalizes the connection considered by Schouten-Dantzig [11] and Chern [1] in the Hermitian case. A systematic
151
comparison of various connections on almost complex manifolds is given in Yano [12]. At this point we want to digress a little to explain vanishing of various components of the torsion form. To the almost Hermitian structure (J, g) we associate the fundamental 2-form (2.8)
$ = ^ ^ 0 ' A 0
J
,
or equivalently $(u,v) — g(u, Jv),
u,vETpM.
When this form $ is closed, (M, J, g) is called an almost Kahler manifold. If, furthermore, J is integrable, then (M, J, g) is a Kahler manifold. Then all these conditions may be tabulated as follows.
d$ = 0
almost Hermitian 0(1,1) = o almost Kahler 0 (1,1) =
0(2,0)
=
integrable J Hermitian 0(1,1)
0
0(0,2)
=
=
0
Kahler 9 = 0
As in (2.5), let il = (fij) be the curvature form of the canonical connection w. In the Hermitian case, the (2,0)- and (0,2)-components of 0, vanish. But in the almost Hermitian case, they may be nontrivial. However, in order to define the holomorphic sectional curvature of an almost Hermitian manifold, we use only the (l,l)-component ft*1-1) = ((nj) ( 1 , 1 ) ) of the curvature. Write (2.9)
= Y,R*iki9k
(n))^
AS
'-
Since we are using unitary frames, we have R
= 2 ^ 9hjRikl
fjkl
=
R
\kV
For a unit vector u = J2 u%tu w e define the holomorphic sectional curvature in the direction spanned by u to be (2.10)
3.
H(u) =
1,1
fl(n<
>(u,«)u,ti) = Y,
Rfjk^jukul.
Almost complex submanifolds
Let M' be an almost complex submanifold of (M,J); at each point pe M' the tangent space TPM' is stable under J. A Hermitian metric g on
152
M induces a Hermitian metric on M'. We shall compare the holomorphic sectional curvature of M' with that of M. Let 2n = d i m M and 2m = dimM'. We choose a local co-frame field 61, • • •, 6n in such a way that 9m+1 = 0, • • •,6 n = 0 on M', (i.e., ei, • • •, e m are tangent to M1). Let u> = (ujj) be the canonical connection on M. We restrict the first structure equation of the canonical connection of M to M'. Since 6r = 0 for r = m + 1, • • •, n, we have d f l ^ - ^ T t d ^ A ^ + e0,
(3.1)
a = l,---,m,
6=1
m
(3.2)
O = d0 r = - ^ c d £ A 0 6 + e r ,
r = m + l,---,n
6=1
While the first equation (3.1) says that u/ = (cd£)ai6=i,...iTO defines the canonical connection on M'. Since 0 has no (1, l)-components, (3.2) implies that, restricted to M', the forms w£, 1 < 6 < m
(3.3)
W = W-
u
b**ra,
Y,
(a,b = l,---,m)-
Since ujra are of degree (1,0) on M', write ro
^=x;^c-
(3.4)
c=l
As in (2.9), set Prom (3.3) and (3.4) we obtain n
(3-5)
Kid = Kid-
E
KM+
r=M+l
So, for any unit tangent vector u = YlT=i u
(3.6)
H'(u) = H(u) -
£ r=m+\
n
IE a,c=l
ft
aC«a«c|2
< H(u).
153
4.
Conformal changes
Now we consider conformal changes of an almost Hermitian manifold (M, J, g) with local unitary co-frame field 91, • • •, 6n and the canonical connection u = (w*). Let / be a real valued function on M, and consider a Hermitian metric h defined by h = e2fg.
(4.1)
Then with respect to a local unitary co-frame field ip1=efel,---,ipn
= ef9n
for h, the canonical connection %/J — (ipj) for (M,J,h) terms of w and / as follows:
(4.2)
^=u)
can be expressed in
+ {d'f-d"f)8),
so that
(4.3)
V = -^VJA<^'+$\
where $* = e ' 0 * + 2 d 7 A p \ A simple calculation gives its curvature \£: (4.4)
* = fi + d{d'f - d"f)I = n -
2(d'd"f)I
and 9&M = H ^ 1 ' -
(4.5)
2(d'd"f)I.
This is significantly simpler than conformal changes of metrics in Riemannian geometry. If v is a unit vector for the metric h = e2fg, then u = e^v is a unit vector for g. Let Hg and Hh be the holomorphic sectional curvatures for g and h, respectively, i.e., Hg(u)=g(rt1<1Hu,u)u,u),
Hh(v)=h(¥1>1Hv,v)v,v).
Then from (4.5) we have Hh{v)
= =
h{rt1'1\v,v)v,v)-2h((d'd"f){v,v)v,v) e- 0(n< >(u,«)u,ti) 2e-2f(d'd"f)(u,u). 2/
lll
Hence, (4.6)
Hh(v) = e- 2 / H f f (u) -
2e-2f{d'd"f)(u,u).
154 5.
Differential geometric criterion for hyperbolicity
One of the sufficient conditions for a complex manifold to be (complete) hyperbolic (in the sense that its intrinsic pseudo-distance is a (complete) distance) is that it admits a (complete) Hermitian metric with holomorphic sectional curvature bounded above by a negative constant. We will show that this can be generalized to the almost complex case. Essential points of the proof are Ahlfors' Schwarz lemma and the fact (shown in (3.6)) that the holomorphic sectional curvature of an almost complex submanifold of M does not exceed that of M. We recall Ahlfors' Schwarz lemma. (5.1) Theorem Let D be the unit disk {\z\ < 1}, and go the Poincare metric, normalized so that its curvature is — 1. Let g be any Hermitian pseudo-metric on D whose curvature is bounded above by —1. Then 9
<9D-
By a pseudo-metric we mean that g may degenerate at some points. At such points, the curvature is not defined. Our assumption is that the curvature is bounded by —1 only at the points where the pseudo-metric is non-degenerate. Prom (3.6) and (5.1) we obtain (5.2) Theorem. Let (M,J,g) be an almost complex manifold such that its holomorphic sectional curvature is bounded above by —1. Then every holomorphic map f:D —> M is distance-decreasing, i.e., f*9 < 9DProof. Set g = f*(g). Since g vanishes where / is degenerate, it suffice to consider the points where / is non-degenerate. Let a 6 D be such a point and U a small neighborhood so that / : U -> M is an imbedding. By (3.6) the curvature of g in U is bounded above by —1. By Theorem (5.1), 9 < 9D on U. D As a consequence we obtain the following differential geometric criterion for hyperbolicity. (5.3) Theorem. Let (M, J,g) be an almost Hermitian manifold. If its holomorphic sectional curvature is bounded above by a negative constant, then (M,J) hyperbolic. If g is moreover complete, (M,J) is complete hyperbolic Proof. By multiplying g with a suitable constant, we may assume that the holomorphic sectional curvature is bounded above by —1. Let p denote the Poincare distance on the unit disk D defined by go- Let Sg be the distance function on M defined by the metric g. Finally, let ^ M be the intrinsice pseudo-distance of (M, J)
155
Let f:D -» M be a holomorphic mapping, and g = f*g. 9 < 9D, which means that the mapping
By (5.2),
f:(D,p)^(M,SB) is distance-decreasing. We recall the following characteristic property of d,M- if 8 is any pseudodistance on M such that every holomorphic mapping f:(D,p) —¥ (M,5) is distance-decreasing, then 6 < dm- By this property of d^, we have 8g < dM-
Since 5g is a distance, dM is also a distance. If Sg is Cauchy-complete, so is dM• Given a point p 0 of an almost Hermitian manifold (M,J,g), there is a positive function / denned in a neighborhood V of p 0 such that the Hermitian metric h = e2fh has holomorphic sectional curvature < — 1. We indicate an explicit construction of such a function / . Let 9 = (01, • • •, 6n) be an adapted unitary co-frame field around p 0 . (By " adapted", we mean that the canonical connection form w = (a;!-) vanishes at po- For the existence of such frame field, see [8]). Take a complex local coordinate system z1, • • •, zn with origin at po such that 8l = dz'
at
po-
The local complex structure given by z1 • • •, zn has nothing to do with the given almost complex structure J (except at the origin po). Let r be a small positive number such that the coordinate system z1, • • •, zn is valid in the ball
c/r = {iNi2 = Ei^i 2 < r 2 }Now, we choose as / the following function defined on Ur:
Then (4.6) and calculation show that if r is sufficiently small, the holomorphic sectional curvature Hh(v) of h at po is negative, (see [8] for details). By continuity, in a small neighborhood V of po, the holomorphic sectional curvature of h remains negatively bounded. In [8] we used a slightly different function / . By (5.3) V is hyperbolic. Thus, (5.4) Theorem. Let (M, J ) be an almost complex manifold. Then every point po £ M has a hyperbolic neighborhood. In [8] I claimed that the neighborhood I constructed is also complete without valid proof. Although I believe that it is complete, the question is
156
still open. In dimension 4, Debalme and Ivashkovich [2] have established local complete hyperbolicity by a different method.
References [I] S-S. Chern, Characterisitc classes of Hermitian manifolds, Ann. of Math. 47 (1946), 85-121. [2] R. Debalme and S. Ivashkovich, Complete hyperbolic neighborhoods in almost complex surfaces, International J. Math., 12 (2001), 211-221. [3] C. Ehresmann, Sur la theorie des espaces fibres, Colloq. Intern'l C.N.R.S. Toplogie algebrique, Paris (1947), 3-35. [4] C. Ehresmann and P. Libermann, Sur les structures presque hermitiennes isotropes, C. R. Acad. Sci. Paris 232 (1951), 1281-1283. [5] M. Gromov, Pseudo-holomorphic curves in symplectic manifolds, Invent, math. 82 (1985), 307-347. [6] R. Hermann, Compact homogeneous almost complex spaces of positive characteristic, Trans. Amer. Math. Soc. 83 (1956), 471-481. [7] S. Kobayashi, Hyperbolic Complex Spaces, Springer-Verlag, 1998. [8] S. Kobayashi, Almost complex manifolds and hyperbolicity, Result. Math. 40 (2001), 246-256. [9] S. Kobayashi and K. Nomizu, Foundations of Differential Geometry, vol. 2, Wiley, 1969. [10] A. Nijenhuis and W. B. Woolf, Three integration problems in almost complex manifolds, Ann. of Math. [II] J. A. Schouten and D. van Danzig, Uber unitare Geometrie, Math. Ann. 103 (1930), 319-346. [12] K. Yano, Differential Geometry on Complex and Almost Complex Spaces, Pergamon Press, 1965.
T h o m Isomorphism of Equivariant Cohomology MEI Xiang Ming Department of Mathematics, Capital Normal University, Beijing 100037, P. R. China Abstract In this paper, we generalizes the Thom isomorphism from ordinary cohomology to equivariant cohomology.
1. Introduction Let M be a compact oriented differential manifold, iV be its closed submanifold, dim M - dim N — r, we have the cohomology exact sequences (see [1], §2) H*-1{M\N)
-4 H*(M,M\N)
3
A
H*(M)
and the Thom isomorphism H*~r{N)
4 H*{M,M\N)
»
ff*(r-LM,7,-LAfy\0
« #£(TxiV),
where T^N denote the normal bundle of N in M, it is diffeomorphism to the e-tubular neighborhood N(e) of N in M, HQ denotes the cohomology with compact support, j» denotes the zero extension in M. In this paper, we generalize the above result to the equivariant cohomology. Let G be a compact Lie group acts left on M, g be its Lie algebra, X £ g generates a vector field X*,N is the zero set of X*. The main results of this paper is to construct a concrete equivariant form with compact support $ e ZcGiT^N) such that: it has the similar properties as Thom form in the ordinary cohomology (see [4], p.32) and the correspondence [OJ] € H*(N) H+ [UJ A $] € H^aiT^N) gives the Thom isomorphism of the equivariant cohomology.
2. Equivariant Cohomology Let M be a compact oriented differential manifold, dim M = n, G is a Lie group acts left on M, G is also acting on C°°(M). For V> € C°°(M), (g-
158
Let g be the Lie algebra of G, VX e 0 generates a C°° vector field X* on M: ( * » ( * ) = ^[4>exp(-eX)
• x)]e=0.
Let A(M) be the algebra of the differential forms of M, the exterior differential d and the contraction operator i(X*) are both antiderivations: it means that, for Va, /? G A(M), + ( - l ) d e s « a A d0,
d(aAP)=daAp
i{X*)(a A P) = i{X*)a A /J + ( - l ) d e g a a A t(Jf*)/9. Let Cx* be the Lie derivative with respect to the vector field X*, it satisfies the homotopy formula: Cx- - d • i{X*) + i(X*)d, so that Lie derivative is a derivation, i.e., for Va,/3 6
A(M),
£x> (a A p) = Cx* (a) A /3 + a A Cx> (/?)• If a e £ ( M ) , Cx'Ot = 0, a is called the equivariant differential form of M, denoted by a £ AG{M), AG{M) is a subalgebra of A(M). Introduces a operator ds on the algebra AG(M) such that: for Va € AG(M), dga = da — i(X*)a. Using the homotopy formula, we can prove that: dea g AG{M) and d^a = 0, for Va G > 4 G ( M ) , so that (^4G(M),dg) constitutes a complex, it is called the equivariant complex of M. This is the de Rham model of the equivariant cohomolgy of M. If a € AG{M) satisfy dga = 0, a is called closed equivariant form; if there exists 0 G AG{M), such that a = dBP, then a is called exact equivariant form. Defines the subalgebra of AG(M) as follows: ZG(M)
= {ae AG(M)
: dsa = 0},
BG{M) = {a G AG(M) : there exists /3 6 .4 G (M), such that a = dB/3}. It is evident that: BG{M) C ZG(M), HG(M) =
then ZG(M)/BG(M)
is called the Equivariant cohomology of M. Let a G . 4 G ( M ) , a'"' be the homogeneous part of degree n in a, we define
f a= f aW.
159
If a £ BG(M), f JM
a=
then a = dgp = dp -
i{X*)p,
f aW = f {dp- i(X*)p)W JM
JM
= f d^gln-i] = 0, JM
because M is compact. So that: if a G ZG(M) the integral fM a depends only the cohomology class [a] € HQ(M).
3. Kobayashi Lemma Let M be a compact oriented differential manifold, G is a compact group acting left on M . Introduce a Riemannian metric h on M, we can always choose that h is G-invariant. Let g be the Lie algebra of G. For VX e a, it generates a Killing vector field X* on M, it means that: For any two vector fields Y and Z on M , we have X* -h(Y,Z)
= h(Cx*Y,Z)
+
h(Y,Cx*Z).
Let N be the zero set of Killing vector field X*, N = \J Ni, where Ni i
is the connected component of N. According to the result of Kobayashi (see [3], p. 60), Ni is the closed geodesic submanifold of M with even codimension, dim Ni = 2(m — r^), (0 < r^ < m). Let T(M) be the tangent bundle of M, the G-invariant metric h of M determines uniquely the Levi-Civita connection V. Let the connection form of V be w, the corresponding curvature form is 0 = du + |[w, u]. The LeviCivita connection V has the following properties: For any two vector fields Y and Z, we have X*-h(Y,Z)
= h(Vx,Y,Z)
+
h(Y,Vx*Z).
Define an operator Ax* — Cx* — Vx* acting on the tangent bundle T(M), it is evident that: h{Ax.Y,Z)
+ h(Y,Ax.Z)
= 0,
so that Ax* is an antisymmetric operator with respect to the metric h. Along the zero set N of X*, X* = 0, so that Ax* \N = Cx* \N.- At the point x € N, Ax* (x) = Cx* (%) is a linear transformation of the tangent space TX(M) of M. Since the Levi-Civita connection is of no torsion, VX*Y - VyX* = [x%Y] = CX*Y, AX*Y = CX*Y - VX*Y =
Vy €
T(T(M)),
-VYX*.
For any vector field Y tangent to Ni, its integral curves belongs to Ni, so that X* = 0 along the integral curves of Y, it shows that: For VY 6
160
r(T(M)), AX*Y = £X*Y
= 0.
So that the action of the operator Ax*(x) on the subspace Tx(Ni) of the tangent space TX(M) is zero. Let T^-(Ni) denote the orthogonal complement of Tx(Ni) in TX(M) with respect to the metric h, dimT^iV-1-) = n — dimTx(Ni) — 2r, (0 < r» < m). Since Ax* is antisymmetric with respect to the metric h, Ax*(x) can be reduced to normal form on T(N) ( Ax.(x)
=
0 -ax
ax 0
\ '
Ax.(x)\ TJ-(Ni) \
0 -ari
ari 0 /
Let T^-Ni be the normal bundle of JV» in M, its fibre is T^{Ni), Vx € Nj. Restricts the Levi-Civita connection V of the tangent bundle T(M) to its subbundle T^iVj), we obtain the induced connection V"1, let the corresponding connection form be OJ-^, and the curvature form is fl1- — duj1- + |[a;-1-,a;-L]. Ax, is an operator acting on the normal bundle, the action of it on every fibre is the above linear transformation Ax, (x). Let Ni(e) be the e-neighborhood of JVj in M. When e is sufficient small, Ni(e) can be coincided to normal ball bundle B±(Ni) on Nt its fibre is a 2rj-dimensional ball on T^Ni with the center x and the radius e. dNi{e) is the boundary of Ni(x), it can be regarded as a normal sphere bundle S±(Ni) on Ni, its fibre is a (2r; - l)-dimensional sphere Si(e) = dBi(e) on T^Ni with center x and the radius e. Define a 1-form 7r on M\N
as follows:
it is evident that i(X*)7r = 0, so that d • i(X*)n = 0. Since dn(X*, Y) = X* • 7r(F) - y • 7T(X*) - TT([X*, F]),
we shall show that: i(X*)dn(Y) = dir(X*,Y) = 0. If F||X*, it is trivial; when YLX*, we can prove that [X*,F]±X, we have also dn(X*,Y) = 0. So that CX' = di(X*)ir + i(X*)dn = 0. It means that n £ AG{M). We can prove further that Cx+dn = 0, so we have also dir € AG(M).
161
Lemma.
lim I
n
=
1
where
Proof. See Kobayashi [4], pp.71-75. Let D be the covariant derivative determined by the Levi-Civita connection, the corresponding curvature form is ft — Dw, we have also the Bianchi identity DU = 0. Generalizes to the equivariant case, we can define the equivariant covariant derivate Dg and the equivariant curvature form fi0 to be
DB = D-i(x*),
ng = n + Ax*.
Similarly, we can prove the equivariant Bianchi identity DgQg = 0 (see [4], 2.8). Let $ € C[g] G , i.e., $ is a complex-valued G-invariant polynomial defined on g, we define
$(ftfl) = *(ft +AxOUsing equivariant Bianchi identity, we can prove that: $(fi g ) € ZQ{M) (see [2], 2.13). $(fi 8 ) is called the equivariant characteristic form of the tangent bundle T(M), [$(ftg)\ G HQ(M) is called the equivariant characteristic class of the tangent bundle. Especially, E(Clg) = ; A „ s/det(ilg) is called the Euler characteristic form of tangent bundle. Now we can rewrite the above Lemma in terms of equivariant terminology. Lemma. e
£ n ™ysi( e )dB 7 r ( i x where -E(n ) is the equivariant Euler characteristic form of the normal sphere bundle dNi(e).
4. The Thorn Isommorphism of the Normal Bundle Let TxNi denotes the normal bundle on iV, in M. B-L(Ni) is the corresponding normal ball bundle, it is coincided to the tubular e-neighborhood Ni(e) of Ni in M via exponential mapping. So we can denote Ni(e) by B±(Ni). The Lie group G keeps the point x E N fixed, since the Lie group
162
G keeps also the Riemannian metric invariant, so it keeps the geodesic passing through the point x and orthogonal to Ni invariant. It shows that: the Lie group G is the fibre-preserving transformation group of BA-{Ni), it is also the fibre-preserving transformation group of the corresponding normal sphere bundle S-^JVj) = dB^Ni) on Nt. Now we shall find the Thom form $ G ZcG(BA-{Ni)) of the normal ball bundle p : B^iNi) -> Ni, where the symbol C denotes compact support, such that the Thom isomorphism: %:HG(Ni)^HCG(B±(Ni)) is determined by the correspondence: [u] i->- [OJ A $]. The following relation of the cohomology in the introduction H—\M\N)
4 H*(M,M\N)
«
HG(TXN)
suggest us that: we must find first a form of M\N, obtain the required Thom form. So that: we put
then acts in by dQ to
$ = d9(p(r)^Ap'£^)), where r is distance from the point of T^-(N) to the origin x,p(r) is a C°° function of the fibre T^-(Ni) « R2ri of the normal bundle T^iNi) which satisfies the condition p(e) = 0, p(0) — 1. The reason to introduce the distance function p(r) is that: it makes the form $ to have compact support. Since G keeps the metric invariant, so that Cx*p{r) = 0, i.e., $ G
ACG(M).
Theorem 1. The differential form <J> is the Thom form of the equivariant cohomology, it has the following properties: (i) * G BCG(B±(Ni))
ZCG(B±(Ni)),
C
(ii) Let ix : B^r(N) -)• B±(Ni)
is the inclusion mapping, then we have
/ .BJ-(JVj) (iii) Let s : Nj ->• B±(Ni)
Proof, (i) It is trivial.
is the zero section, then we have
163
(ii)
[d(p(r)-^-)[2r'"ll|AJE(n^),
= lim/
where Bx(r) is the ball with the center x and the radius r. Since
so that
according the Lemma due to Kobayashi, we have JB±(Ni)
(iii)
s^
•*=mr)*m)=1'
=
^l"a
s*de[p(r)J^)As*p*E(nj)
= <*<»£) *E<#) = *,(£.) A Eft) = E(^).
The theorem is proved. T h e o r e m 2. Let UJ G Zo{Ni), then the correspondence [u] H-> [w A $] determines the Thom isomorphism of the equivariant cohomology
Z:
H^Ni)-+HZaiB^Ni)).
Proof. According to the Theorem 1 (ii)
/
p*uA$=
JBi-(Ni)
L A | JNi
L
i * $ ] = / u>,
JB^(Ni)
J
./JV;
since the above integral depends only the equivariant cohomology class, so we have the homomorphism of the equivariant cohomology 1: H'G{Ni) -> H*CG{B^{Ni)), Let w 6 ZcGiB^iNi)), ±
of Ni in B (Ni)
[u] H> [U A $].
dgu) — 0, and ATj(r) is the r-tubular neighborhood
« Wi(e), then in the B-1 (Ni)\Ni(r),
*(d^ A 4
we have
164
Notes that the zero section s(iVj) of BL(Ni) is the deformation retract of Bx(Ni),s • p ~ identity, so that w = p*s*uj + dr, but u has compact w support |nR±fjy.i = 0 ) t n e n w e have /
tj = lim /
d(-—Aw
r
-*0JBJ-(Ni)-Ni(T)
JB-^(Ni)
= I
B
-^(P*S*0J
JdB-^(Ni)
since fdN.,r\
Kd n
'
+ dT),
"gir
dr = 0, according to the Kobayashi Lemma, we have
JBHNi)U~ Jt*i^ ^ JSi (r)dgTt) S " ~ JN. E(Q^ ) ' then we obtain the another equivariant cohomology homomorphism p. : NZoiBHNi))
~> flc(JVf), M -> [ ^ y ] •
Now we shall that: the homomorphism T and p are mutually inverse, so that they are both isomorphisms. If a; € Za(Ni), according to the Theorem l(iii),
P. • T H = Pt[p* A $] = [
P E{n±)
\ = M.
If w € ZcG(Ni), notes that s(iVj) is the deformation retract of so that [td] = [p*s*w], then we have
„-
r T
^r
s*u
1
r
p*s*w
BJ-(Ni),
i
= [pVo.] A [d, ( / » M ^ ) ] = H A [1] = H , the Theorem is proved.
References 1. Atiyah, M. &; Bott, R., The moment map and equivariant cohomology, Topology, 23(1984), 1-28. 2. Berline, N. & Vergne, M., Zeros d'un champe de vecteur et classes characteristique d'equivariante, Duke Math. J., 50(1983), 537-549.
165
3. Berline, N., Getzer, E. & Vergne, M., Heat kernel and Dirac operators, Math. Wis., Vol. 298, Springer-Verlag, 1992. 4. Kobayashi, The transformation group in differential geometry, SpringerVerlag, 1972. 5. Bott, R. & Tu, L., Differential forms in algebra topology, SpringerVerlag, 1982.
Horizontally conformal F-harmonic maps * XIAOHUAN MO AND CHUNHONG YANG Key Laboratory of Pure and Applied Mathematics School of Mathematical Sciences Peking University, Beijing 100871, China E-mail: [email protected] and [email protected]
Abstract F-harmonic maps are maps which extremize the F-energy functional.
In this
article we study the F-harmonicity of horizontally weakly conformal maps and the conformal properties of F-harmonic maps. Key words and phrases: F-harmonic map, horizontally weakly conformal maps, F-tension field 1991 Mathematics Subject Classification: 58E20.
1
Introduction
Harmonic maps introduced by Eells and Sampson in 1964[5], are critical points of the 'Dirichlet' energy functional
\ f \\d
and in this respect can be seen as a generalisation of geodesies. Harmonic maps are solutions to an elliptic system of partial differential equations, which in general is non-linear. Harmonic maps are very important both in classical and modern differential geometry. 'This work is supported by the National Natural Science Foundation of China 10171002
166
167
A particularly interesting subclass of harmonic maps is horizontally weakly conformal harmonic maps. The interplay between the analytical condition (harmonicity) and the geometrical one (horizontal weak conformality) is a rich source of properties. Baird and Eells [2] studied some conformal properties of harmonic maps. They showed that if the dimension of target is equal two, then the fibers of horizontally weakly conformal harmonic maps are minimal submanifolds in the domain manifold. When the codomain is not of dimension two, R. Pantilie showed that the harmonicity for a horizontally conformal map is characterized by the property that the parallel displacement defined by the horizontal distribution (i.e. the distribution orthogonal to the fibres), preserves the mass of fibres[9,10]. This generalizes the well-known fact that a Riemannian submersion is harmonic if and only if the parallel displacement defined by the horizontal distribution preserves volumes. Viewing harmonic maps as maps which extremize the energy functional, their most natural generalization is the notion of F-harmonic maps, i.e. maps which extremize the F-energy functional[1]. They encompass p-harmonic maps [3,12] and exponentially harmonic maps[4]. In this paper we discuss the F-harmonicity of a horizontally weakly conformal harmonic map. In particular, we show the following Theorem 5 A horizontally conformal submersion with dilation A is an F-harmonic map if and only if the parallel displacement defined by the horizontal distribution preserves the mass of the fibres, where the fibres are given the mass density A 2 _ n .F'(§A 2 ). When F(t) = t and (2i) 2 /p, this theorem has been proved in [9,10] and [7], respectively. We also extend the results of conformal properties due to Baird and Eells, Baird and Gudmundsson, and Takeuchi.
168
2
T h e Fundamental Equation for a Horizontally Conformal M a p
Let F : [0, oo) ->• [0, oo) is a C 2 function such F' > 0 on (0, oo). For a smooth map <> / : (M, g) ->• (N, h) between Riemannian manifolds (M, g) and (N, h), we define the F-energy of <j> by ,|2-
L'ffl
dv
where \d(j>\ denotes the Hilbert-Schmidt norm of the differential d
> 4), (l + 2t) a (a > l.dimM =
2) and exp t, respectively. We shall say that <j> is an F-harmonic map if it is a critical point of the F-energy functional, which is a generalization of harmonic maps, p-harmonic maps or exponentially harmonic maps. Recall that a map 4> : M -¥ N between Riemannian manifolds is Fharmonic if and only if its F-tension field TF{<J>) := F' 0 ^ \
r+V
(F'
( ^ ) )
vanishes. The reader is referred to [1] for a detailed account of F-harmonic maps. Call a smooth map (j> : M -» N between Riemannian manifolds horizontally (weakly) conformal if for any point p € M which is not contained in the critical set C^ = {p 6 M\d(j)p = 0} of
>M= Ofor all Y € Kercty,}
of Ker d
or, locally,
<> / satisfies the following equations[13]
*
dxl dxi
v
'
for some function A : M -> [0, oo) called the dilation of
169
Proposition lLet <j> : (M, g) -> (N, h) be a horizontally conformal map with dilation X. Let rp be the tension field of <j) and r] the trace of the second fundamental form of the vertical subbundle of
(2)
where r is the tension field of (j>. In general, if / : R -> R, and h: M -* R then V ( / o h) = /'V/i
(3)
Note that 0 is a horizontally conformal map with dilation A, i.e. (jfh = ^29\n where % — (Kerdcf))1 is the horizontal distribution of <j>. Then we have \d(j>\2 = nX2
(4)
On the other hand, the F-tension field Tp of
- ' '(^)-v(r(M)) Substitute (3) and (4) into (5) and use (2) we have
T
" = -5F t(" -2)F' (? A! ) " " ^ &')}VA!"
F
' (1 A ! )' <6)
Since F' > 0 on (0, oo), (6) is equivalent to 2
TF
^FTW)+%
{n-2)F'^X
)-nX2F"(^)r,
0
(7)
170
By a straightforward computation, we obtain
vln
nfA^)-
Fif^j
vlnA
(8)
Substitute (8) into (7) we have (1). Notice that for F(t) — t we obviously recover the fundamental equation in terms of the tension field T.
3
An Extension of Baird and Eells' Result
Our fundamental equation (1) implies the following T h e o r e m 2 Let (Mm, g) and (Nn, h) be Riemannian manifolds and
Then two of the
following conditions imply the other. (a) <j> is an F-harmonic
map,
(b) (f> has minimal fibres; F^xi)
*s vertical.
This result can be viewed as an extension of Baird-Eells' observation that, if the dimension of target is equal two, then the fibers of horizontally weakly conformal harmonic maps are minimal submanifolds in the domain manifold, and when the codomain is not of dimension two, then the fibers of horizontally homothetic harmonic maps are minimal submanifolds in the domain manifold. Recall that a horizontally conformal map <j> : M ->• N is said to be horizontally homothetic if %V(1/A 2 ) = 0 on MQ. AS an immediate consequence of Theorem 2 we have Corollary 3[3,12]Le£ ( M m , g) and {Nn, h) be Riemannian and (j> : M —> N be a horizontally conformal submersion. harmonic if and only if the fibres of <j> are minimal in M.
manifolds Then <j> is n-
171
4
F-harmonicity of Horizontally Conformal Maps
Let <j> : (M, g) -> (N, h) be a Riemannian submersion. Ara showed that the fibers of (f> are minimal submanifolds if and only if <j> is F-harmonictl]. Hence we have the following: Proposition 4 A Riemannian submersion is F-harmonic if and only if the parallel displacement defined by the horizontal distribution preserves volumes. In the general case we have Theorem 5 A horizontally conformal submersion with dilation A is an F-harmonic map if and only if the parallel displacement defined by the horizontal distribution preserves the mass of the fibres, where the fibres are given the mass density A 2 _ n F'(f A2). Proof We shall denote by V the projection on the vertical subbundle. Put /* := A 2 - F ' ( | A 2 ) Then, for any basic vector field X on M we have V(Cx(jiu))
=
V*[(Xn)u)+n£xu]
=
(X(j,)V*LJ +
=
{X(J,)UJ -
=
(Xp)u - ng (X,
= =
fiV*(Cx^)
fj,g(X, r))u> 2 n
F,(
(Xfi)u) + X ~ g(X,
r
| ^ ) + V In F^lt))
TF)UJ
u
- fig(X, V In fj,)w
2 n
X - g(X,TF)iJ
Remark When F(t) = t, we reduce R. Pantilie's result[9,10]. Acknowledgements.
I would like to express my special thanks to Pro-
fessor J. C. Wood for his constant encouragement during the preparation
172
of this work. I would also like to thank Professor S. Montaldo for his very helpful suggestions.
References [1] M.Ara, Geometry of F-harmonic maps, Kodai Math. J., 22(1999), 243263. [2] P.Baird and J.Eells, A conservation low for harmonic maps, In Geometry Symposium Utrech 1980, Lecture Notes in Math. 894 (SpringerVerlag,1981) pp.1-25. [3] P.Baird and S.Gudmundsson, p-harmonic maps and minimal submanifolds, Math. Ann. 294(1992), 611-624. [4] L.Cheung and P.Leung, The second variation formula for exponentially harmonic maps, Bull. Austral. Math. Soc. 59(1999), 509-514. [5] J.Eells and J.H.Sampson, Harmonic mappings of Riemannian
mani-
folds, Amer. J. Math. 86(1964), 109-160. [6] S.Gudmundsson, On the existence of harmonic morphisms from symmetric spaces of rank one, Manuscripta Math. 93(1997), 421-433. [7] H.Jin and X.Mo, On submersive p-harmonic morphisms and their stability, accepted for publication in Contemporary Mathematics. [8] X.Mo, Horizontally conformal maps and harmonic morphisms, J. Chin. Cont. Math., 17(1996), 245-252. [9] R. Pantilie, Harmonic morphisms with one-dimensional fibres, Intern. J. Math., 10(1999), 457-501. [10] R. Pantilie, On submersive harmonic morphisms, In Harmonic morphisms, harmonic maps and related topics, (eds. C.K.Anand, P.Baird, E.Loubeau and J.C.Wood). Brest 1997, Pitman Research Notes in Mathematics, CRC Press (1999), 23-29.
173
[11] J. Sacks and K. Uhlenbeck, The existence of minimal immersions of 2-spheres, Ann. of Math., 113(1981), 1-24. [12] H. Takeuchi, Some conformal properties of p-harmonic maps and a regularity for sphere-valued p-harmonic maps, J. Math. Soc. Japan 46 (1994), 217-234. [13] J.C.Wood, The geometry of harmonic maps and morphisms, Preprint of University of Leeds, 34 (1999).
H A R M O N I C M A P S B E T W E E N C A R N O T SPACES
S E I K I NISHIKAWA Mathematical
Institute, E-mail:
Tohoku University, Sendai 980-8578, [email protected]
Japan
A fc-step Carnot space is defined to be a homogeneous Riemannian manifold of negative curvature obtained as a one-dimensional solvable extension of a certain graded nilpotent Lie group, called a fc-step Carnot group. This class of manifolds includes the rank-one symmetric spaces of non-compact type as examples. We discuss the asymptotic behavior of proper harmonic maps from a fc-step Carnot space to an (-step Carnot space, where k ^ I in general.
1. Introduction 1.1. Harmonic
Maps
Let M = (M,g) and M' = {M1 ,g') be Riemannian manifolds of dimension m and m', respectively, and u : M -> M' a C 2 map from M to M'. For the differential du : TM -> TM1 of u, the covariant derivative Vdu of du is defined by Vdu(X,Y)
= Vux~lTM'du(Y)
-
du(VlMY),
where V ™ and Vu ™ denote the Levi-Civita connection of M and the pull-back of the Levi-Civita connection of M' by u, and X,Y £ T(TM) are vector fields on M, respectively. Then the tension field T(U) of u is defined to be the divergence of du, that is, 771
T(U) = Trace,,Vdu = ^
Vdu(Ei,
E{),
«=i
where {Ei} are orthonormal frame fields on M. We call u a harmonic map if the tension filed r(u) of u vanishes identically. In terms of local coordinate systems (xl) and (a;'a) of M and M', the Riemannian metrics g and g' are given locally by 777-'
777
9 = Yldijdx^x^
g1 = Y^
«=1
a,P=l
174
g'a0dx'adx'0,
175
and u is written as u{x) = {ul{x\...,xm),...,um\x\...,xm))
=
{u«{xi)).
Then the energy density e(u) of u is defined to be the squared norm of du, that is,
e(u) = \du\2=£
£>V«*(«)»£g!,
i , j = l ct,/3=l
and for a relatively compact domain fl on M the energy .EQ (U) of u over fi is defined by EQ{U)
= s/»«<«'
= - I e{u)dng,
where \xg denotes the canonical measure on M given by the Riemannian metric g. Then the energy EQ(U) of u defines a functional EQ : C2(M, M1) -> R on the space C2(M, M') of C 2 maps from M to M', and the harmonic map equation T(U) = 0 is nothing but the Euler-Lagrange equation for the energy functional, which is expressed in local coordinate systems as
A«a+EE^»H=°.
a = l,...,m', (1.1)
» J = 1 /3,7=1
where A is the Laplace-Beltrami operator on M and T'g (u) denote the Christoffel symbols of the pull-back connection V" ™ . Since the harmonic map equation (1.1) is a system of semi-linear elliptic partial differential equations of second order, any C2 harmonic map u is indeed a smooth map. 1.2. Hadamard
Manifolds
A complete, simply connected, connected Riemannian manifold M of nonpositive curvature is called a Hadamard manifold. It is well-known by the Cartan-Hadamard theorem that M is diffeomorphic to the Euclidean space of the same dimension, and there is a unique geodesic joining any two points of M. We say that two unit speed geodesic rays 7 and a in M are asymptotic if there exists a positive constant C such that the distance d{^{t),a{t)) of 7(f) and a(t) in M satisfies d(j(t),a(t)) < C for all t > 0, and define a point at infinity of M to be an equivalence class of asymptotic geodesic rays in M.
176
Let M be a Hadamard manifold of dimension m and M(oo) denote the set of points at infinity of M. If we set M = M U M(oo), then it is known that M(oo) is homeomorphic to the (m — l)-dimensional sphere 5 m _ 1 and M to the m-dimensional closed ball Bm, relative to a natural topology, called the cone topology, of M which induces on M the original topology of M. We call M the geometric compactification or the EberleinO'Neill compactification of M and M(oo) the idea/ boundary or the sphere at infinity of M ([*]). Example 1.1. (the real hyperbolic plane) TTie most familiar example of a Hadamard manifold is the real hyperbolic plane RH2, the ball model of which is defined to be the open unit disk equipped with the Poincare metric gp of constant negative curvature —1, that is, RH2 = f{z G C | |*| < 1}, gP
4\dz\
(i-N
2N2
In this model, two geodesic rays in M = RH2 are asymptotic if and only if they meet the boundary circle S1 = {z £ C \ \z\ = 1} at the same point, and hence the ideal boundary M(oo) can be identified with Sl. Note that the Poincare metric gp blows up at the boundary circle S1 as \z\ —> 1 so that the corresponding Laplace-Beltrami operator A is degenerate everywhere on the ideal boundary Sl. The circumstance can be seen more clearly if we take the upper half plane model H+ of the real hyperbolic plane RH2, that is, 9
(
rr 2 r H +=({z
i—r
= x + V^ly
This model is isometric to RH2 RH2 given by
^i
„-,
G C | y > 0}, gH =
dx2 + dy2
^ -
under the Cayley transform $ : H^
$( Z ) = v ^ i i - X — ,
zeHl,
r2
(1.2)
which maps the real axis {z e C \ y = 0} into the boundary circle S1. Therefore we may regard $ as defining a coordinate neighborhood of RH2 around its ideal boundary S1. Hence, by taking a family of these Cayley transforms as a system of coordinate neighborhoods at the ideal boundary, the geometric compactification M = RH2 U S 1 of the real hyperbolic plane RH2 admits a structure of smooth manifold with boundary, relative to which M is diffeomorphic to {z £ C \ \z\ < 1}. Then the Poincare metric gH of the upper half plane model H2, yields a local coordinate representation of
177 the Poincare metric gp of RH2 given by $ , that is, (
with respect to a coordinate
4\dz\2
\ _
* V(l-|z|»)V-
dx2+dy2 2
y
neighborhood
'
which shows more exactly how the metric on the real hyperbolic plane
{1 d)
-
RH2
blows up near the ideal boundary as y -> 0.
1.3. Proper
Harmonic
Maps
Let M = ( M , g) and M' = (M1, g') be H a d a m a r d manifolds of dimension m and m ' , respectively, and M = M U M(oo) and W = M'U M'(oo) be their geometric compactihcations. Let u : M —> M' be a proper harmonic m a p from M t o M'. Here the properness of u means t h a t if {pj} is a sequence of points in M diverging to the ideal boundary M(oo) of M, then {u(pj)} also diverges in M' to the ideal boundary Af'(oo) of M'. Suppose now t h a t u extends to a continuous m a p u : M —> M' between the geometric compactihcations. Since u is assumed proper, it follows t h a t u maps the ideal boundary M(oo) into the ideal boundary M ' ( o o ) , and thus we have the boundary value / = it|M(oo) : M(oo) -> M'(oo) of u between the ideal boundaries. Our primary objective is to investigate to what extent the asymptotic behavior of a proper harmonic m a p u near the ideal b o u n d a r y is determined by its boundary value / . T h e first remarkable result in this regard was obtained by Li and Tarn [4] in the case where M and M' are real hyperbolic spaces RHm and RHm . Namely, identifying the ideal boundaries of RHm and RHm with Sm~l and Sm - 1 , respectively, they proved the following rigidity at infinity for proper harmonic maps with C 1 boundary value. T h e o r e m 1.1. /*/ Let u be a proper harmonic map from RHm to RHm , which extends to a C1 map between the geometric compactifications. In the ball model, in terms of the polar coordinates (p,rj) = (p, n1,. .. ,77 m _ 1 ) on Bm and (r,8) = ( r , 8 1 , . . . ,6m _ 1 ) on Bm , express u as u{p,rj) = (r(p,rj),9(p,rj)). Let f : 5 m _ 1 —> Sm ~x be the boundary value of u, and assume that the energy density e(f) of f, relative to the canonical metrics on S"1^1 and Sm ~1, is nowhere vanishing. Then, on 5 m _ 1 , u must satisfy
£ = J«L,
£-0,
op
dp
Vm- 1
.<«<.M
d.4)
178
It follows from (1.4) that the expression u(p,r]) = (r{p,rj),6(p,r])) has the following expansion near the ideal boundary 5 m _ 1 as p —> 1: r(p,r,) = 1 - (1 - p)v/e(f)(r1)/(m-l)+ f(M)=f(i))+o(l-p),
o(l - p),
l
Prom this expansion, Li and Tarn proved the uniqueness of proper harmonic maps with C1 boundary value. To be precise, let u and v be proper harmonic maps from RHm to RHm , which extend to C 1 maps between the geometric compactifications. If u and v have the same boundary value / on the ideal boundary S™ - 1 and the energy density of / : S"1"1 ->• Sm ~l is nowhere vanishing, then u and v are identical. It should be remarked, however, that for proper harmonic maps u : RHm ->• RHm' which are not C 1 up to the ideal boundary Sm~1, we do not in general have the uniqueness. Indeed, Li and Tarn [4] constructed a two-parameter family of harmonic diffeomorphisms of RH2 that are only 1/2 Holder continuous at a boundary point and smooth everywhere else, and whose boundary value is the identity map of S 1 . 2. C a r n o t Spaces 2.1. Homogeneous
Manifolds
of Negative
Curvature
In order to study the asymptotic behavior of proper harmonic maps, we now review briefly the geometry of homogeneous Hadamard manifolds. Let M = (M, g) be a Hadamard manifold of dimension m with the ideal boundary M(oo), and suppose M to be homogeneous, that is, the isometry group I{M,g) of M acts transitively on M. Then it is known that there exists a solvable subgroup G of Jo (M, g), the identity component of I(M, g), that acts simply transitively on M ([3]). Therefore we can identify M with a solvable Lie group G and the Riemannian metric g on M with a left invariant metric ( , ) on G. Consequently, any homogeneous Hadamard manifold M can be represented as a simply connected solvable Lie group G endowed with a left invariant metric ( , ) of nonpositive curvature. For instance, if M is a rank one symmetric space of noncompact type, then I0(M,g) has the Iwasawa decomposition N • A • K, where N is a nilpotent subgroup, A is an abelian subgroup and K is a maximal compact subgroup, respectively. Since K is known to be the isotropy subgroup of a point in M, by setting G — N • A, the semi-direct product of N with A, we obtain a solvable Lie group G acting simply transitively on M. If we further assume that M is of strictly negative curvature, then it follows that G is a one-dimensional solvable extension of a nilpotent Lie
179 group TV. Indeed, let g denote the Lie algebra of G and n = [Q,B] be its derived algebra. Then we know, since g is solvable, that n is a nilpotent subalgebra of g. Moreover, it follows from the curvature condition that the orthogonal complement n x of n must be one-dimensional, that is, n x = R{H} with a choice of a generator H (see [3] for details). Since G is a simply connected solvable Lie group, this implies that G is a semi-direct product TV xi i? of a nilpotent Lie group TV with the abelian group R, where TV = [G, G] is the commutator subgroup of G, which is a simply connected subgroup with Lie algebra n. As a result, G is diffeomorphic to the product manifold TV x R of TV and the real line R, and hence, identifying R with the positive half line R+ = {y € R \ y > 0} by the exponential function R3 s >-> y = es £ R+, we obtain a diffeomorphism * :TV x R+ 9 (n,y) ^-n-logy
€ G = N x R,
(2.1)
which identifies G with the half space TV x R+, the product manifold of TV and R+. Consequently, a homogeneous Hadamard manifold M of strictly negative curvature can be represented as the half space TV x R+ of a simply connected nilpotent Lie group TV with the positive half line R+, which we call a half space model of M. Moreover, we may regard the diffeomorphism $ as a generalized Cayley transform. Indeed, if M is the real hyperbolic plane RH2, then TV is isomorphic to the abelian group R, so that the product manifold TV x R+ is identified with the upper half plane model H2+ of RH2, and the diffeomorphism $ in (2.1), composed with the diffeomorphism given by the simply transitive action of G on RH2, is identical with the Cayley transform $ : H\ -> RH2 in (1.2). 2.2. Rank
One Symmetric
Spaces of Noncompact
Type
In the case of a rank one symmetric space of noncompact type, we see that the nilpotent group TV appearing in the half space model is very restricted. To be precise, let M be a rank one symmetric space of noncompact type, or equivalently a connected symmetric Riemannian manifold of strictly negative curvature, and identify M with a solvable Lie group G with left invariant metric. Recall that the curvature tensor field R of M is parallel with respect to the Levi-Civita connection V of M, that is, VR = 0. Then it follows from this curvature condition that the nilpotent group TV in the half space model TV x R+ of M must be a 2-step nilpotent group at the most. Namely, the Lie algebra n of TV satisfies
[n,[n,n]] = {0}.
180
Consequently, if we set ri2 = [n, n] and ni to be the orthogonal complement of ri2 in n, then the Lie algebra g of G is decomposed into g = n1+n2
+
R{H},
where H denotes a generator of n1-, and n — ni + rt2 becomes a graded Lie algebra. Indeed, letting it; = {0} for i > 3, we have K.rtj] C ni+j,
l
Moreover, we see that ni and n2 are in fact the eigenspaces of the adjoint representation adif of H on n with eigenvalues A and 2A for some nonzero constant A £ R, respectively: m = {X <En\a,dH{X) = i\},
i = 1,2,
and each rank one symmetric space of noncompact type is distinguished by the dimension of ri2, which is equal to 0,1,3 or 7 corresponding to real hyperbolic spaces RHm, complex hyperbolic spaces CHm, quaternion hyperbolic spaces HHm, or the Cayley hyperbolic plane C&H2, respectively. For details, we refer the reader to Heintze [3]. Example 2.1. (complex hyperbolic spaces) Note that, in this context, real hyperbolic spaces are exceptional in the sense that n is abelian, that is, 1-step nilpotent so that n = ni and ri2 = {0}. The nature of the geometry in the case where n is 2-step nilpotent becomes apparent once we look at complex hyperbolic spaces. More precisely, let M be the complex hyperbolic space CHm of dimension m, which is defined to be the open unit ball B2m in the complex Euclidean space Cm of complex dimension m equipped with the Bergman metric QB of constant holomorphic sectional curvature —1, that is,
CH~ = L e c- 11.| < 1}, „ = l ^ ' - ' ^ - ' ^ W ) . As in the case of real hyperbolic spaces, in this model the ideal boundary M(oo) of M is identified in a natural way with the boundary sphere S2m~1 of the ball B2m. Also, it is well-known that the special unitary group 5(7(1, m) of signature (l,m) acts transitively on B2m as isometries, and the isotropy subgroup K at the center of B2m is isomorphic to U(m), which is a maximal compact subgroup in the Iwasawa decomposition of 5(7(1,m). Indeed, the decomposition is given as 5(7(1, m) = N • A • K, where N is a 2-step nilpotent subgroup, called the Heisenberg group, and A is a onedimensional abelian subgroup. By setting G = N x A, we then obtain a
181
solvable subgroup G of SU(l,m), which acts simply transitively on B2m as isometries, so that CHm is identified with a solvable Lie group G with a left invariant metric ( , ) . Let N x R+ be the half space model of CHm obtained from the above identification, and $ : TV x R+ —• CHm the corresponding generalized Cayley transform. Let n = ni + n2 be the graded Lie algebra decomposition of the Lie algebra n of N. Then, under the left translations of N, ni and ri2, being identified with subspaces of the tangent space of N at the identity, define complementary distributions on N. Furthermore, under the generalized Cayley transform $, the ideal boundary M(oo) = 5 2 m _ 1 of CHm is identified with the one-point compactification of the Heisenberg group N, and if we consider the Hopf fibration S2m~l -»• CP171"1 of S2m~l over the complex projective space CPm~l of dimension m — 1, then the distribution 112 corresponds to the vertical subspaces along the fiber and ni to the distribution of horizontal subspaces defined by the canonical contact structure on the odd-dimensional sphere 5 2 m _ 1 . See j5] for more detail. 2.3. k-step
Carnot
Spaces
Motivated by these observations, we now consider a more general class of homogeneous Riemannian manifolds of negative curvature that arises as a one-dimensional solvable extension of certain k-step nilpotent Lie groups, called Carnot groups. More precisely, let G be a simply connected, connected solvable Lie group satisfying the following conditions: (1) G is a semi-direct product of a nilpotent Lie group N and the abelian group R. (2) If n and g = n + R{H} denote the Lie algebras of N and G, respectively, then n has a decomposition n = J2i=i ni m t o ^ subspaces given by ^ = {X e n I adH(X) = i\X},
i = l,...,k,
(2.2)
for some nonzero constant X £ R. Note that, since the adjoint representation adif is a derivation of n, the above decomposition of n defines a graded Lie algebra structure of n, that is, [ni,r\j]Cni+j,
l
(2.3)
with the convention n* = {0} for i > k. In particular, n is a fc-step nilpotent
182
Lie algebra, that is, the k-th derived algebra of n vanishes: [n,---[n,[n,n]]---] = {0}. Moreover, since the eigenvalues of adH are nonzero real numbers, it follows from a theorem of Heintzef3] that G admits a left invariant metric g of strictly negative curvature. Consequently, we obtain a homogeneous Riemannian manifold M = (G, g) of strictly negative curvature. With these understood, we give the following definition. Definition 2.1. A homogeneous Riemannian manifold M — (G,g) of negative curvature obtained as above is called a k-step Carnot space. For example, real hyperbolic spaces are 1-step Carnot spaces, and complex or quaternion hyperbolic spaces and the Cayley hyperbolic plane are 2-step Carnot spaces. Also, it should be noted that if k > 2, then a /c-step Carnot space is not symmetric (cf. Subsection 2.2). Remark 2.1. As our convention in the following, for a given k-step Carnot space M = {G,g), we always choose H, the generator of the Lie algebra of the abelian group R, in such a way that it is of unit length, that is, g(H,H) = 1, and with respect it the eigenvalues iX of the adjoint representation adH are all positive, that is, A > 0. Now, let M = {G,g) be a &-step Carnot space, where G is a simply connected, connected solvable Lie group which is a semi-direct product N x R of a nilpotent Lie group N with R, and g is a left invariant metric on G of strictly negative curvature. Recall that, as seen in Subsection 2.1, M is represented as the product manifold N x R+ of N and the positive half line J R + via a generalized Cayley transform $ : TV x R+ —¥ G given by $ : N x R+ 9 (n, y) i-> n • exp sH € G = N x R,
(2.4)
where s = logy and H is chosen as in Remark 2.1. This half space model N x R+ of M is convenient to study the geometry of M at infinity. In fact, since M is a Hadamard manifold, we have the ideal boundary M(oo) and the geometric compactification M — MuM(oo) of M, respectively. Then, regarding a generalized Cayley transform $ as defining a coordinate neighborhood around the ideal boundary M(oo), we obtain the following proposition which describes the behavior of g near M(oo). Proposition 2.1. f'] Under a generalized Cayley transform $ : N x R+ —> G, the metric g is expressed on a half space model N x R+ of M as a k-ply
183
warped product metric 1
1 + ^9n2+---
* 9=^9n1
1 +^ 9 n
k
dy2 + ^ ,
, s (2.5)
where y denotes the coordinate on R+! A is given in (2.2) and gni + gn2 + • • • + gnk is a left invariant metric on N. As a consequence, we see that, on the half space model N x R+ of M, the eigenvalues {i\ | i = 1 , . . . , k} of the adjoint representation adiJ in (2.2) indeed reflect on the growth order of the metric g as y —> 0. Example 2.2. (1) In the case of the real hyperbolic space RHm, as seen in Subsection 2.2, N is abelian so that k = 1, and with respect to the canonical choice of H, we see that A = 1 and the Poincare metric gp is expressed as
$*gp = ±[(dxl)2 + --. + (dxm-iy] + d^, y y where (x1,... ,xm~l ,y) is the canonical coordinate system of the upper half space model Rm~l x JR+ =* N x R+ of RHm (cf. (1.3);. (2) In the case of the complex hyperbolic space CHm, as seen in Example 2.1, N is the Heisenberg group so that k = 2, and with respect to the canonical choice of H, we see that A = 1/2 and the Bergman metric gs is expressed as .. m~ 1 y
j=l
.. y
m—1 1=1
, 2 y
where (z1,..., zm~l, t) is the canonical coordinate system of the Heisenberg group N £ Cm~l x R, and dzl and dt + ^,7=^ V^l{zidzi - zidzi)/2 give rise to left invariant 1-forms on N if"]). It is immediate from (2.5) that on a half space model N x R+ of M, the unit vector field X = y(d/dy) along R+ directions {(n,y) \ n = const} satisfies VxX = 0 and hence defines geodesic rays which are asymptotic each other. Also, since M is of strictly negative curvature, we see that any two distinct points in M(oo) can be joined by a unique geodesic of M Q1]). As a consequence, the ideal boundary M(oo) of M is identified with the one-point compactification of the Carnot group N. Namely, we have the following Proposition 2.2. On a half space model N x R+ of a k-step Carnot space M, R+ directions define asymptotic geodesic rays in M and hence gives rise to a point at infinity oo € M(oo). Moreover, M(oo) \ {oo} is naturally identified with N x {0}.
184
2.4. Smooth Structures
on
M(oo)
It should be remarked that in the definition of a generalized Cayley transform $ : N x R+ ->• G given in (2.1) and (2.4), we identify the real line R with the positive half line R+ by the exponential function R 3 s ^ y = esG R+. However, for a given fc-step Carnot space M = (G,g), the choice of a diffeomorphism between R and R+ is not unique, and in general it is not apparent a priori which choice should be canonical or expedient. In fact, for a nonzero constant a 6 it, if we choose the exponential function R 3 s H-> y — eas € -R+ in (2.4), then the pullback metric <&*g in Proposition 2.1 takes the form 1 **9=^9n1
1 + ^gn2
1 1 dy2 + --- + ^j^;gnk + ^ ^ ,
(2.6)
and hence the metric g blows up at the ideal boundary M(oo) of M with different growth orders from those in (2.5). Moreover, when taking the corresponding generalized Cayley transforms $ : N x R+ 4 G as local coordinate neighborhoods around the ideal boundary M(oo), these different choices of exponential functions identifying R with R+ define different smooth structures on N x R+ which are not diffeomorphic at the ideal boundary N x {0}, since the coordinate change from y = es to y = eas is not diffeomorphic at y = 0 unless a — 1. Remark 2.2. For a given nonzero constant a £ R, if we take A/a in place of X in (2.2), then it is evident from Proposition 2.1 that on the half space model N x R+ defined by the generalized Cayley transform (2.4), we obtain the pull-back metric $*g that blows up at the ideal boundary M(oo) with the same growth orders as those for the metric in (2.6). Since different choices of the eigenvalues of the adjoint representation adH imposed in the condition (2.2) of the Carnot group G give rise to different smooth structures at the ideal boundary N x {0} on the half space model N x R+ of the Carnot space M = {G,g), it also affects considerably the asymptotic behavior of proper harmonic maps between Carnot spaces. Note here that these different smooth structures are not diffeomorphic each other only on the boundary N x {0}. Before going into details in the next section, we first remark the following typical phenomenon. We now take M to be the complex hyperbolic space CHm of dimension m and M' to be the real hyperbolic space RHm of dimension m', and let M = {G,g) and M' - {G',g') be the representations of M and M' as Carnot spaces, respectively. Recall that the Lie algebras g and g' of G and
185
G' are decomposed respectively as g = ni+n2
+ R{H},
g' = n[ +
R{H'},
and suppose on their half space models N x R+ and N' x R+ that the pull-back metrics $*g and $'*g' are expressed as 1
1
**9 = ^
A
+^
2
d
1
m
"^
A
a
'2
+ f, *V = ± £ (
according to the choices of generalized Cayley transforms $ : N x R+ -> G and $ ' : AT' x U + -» G', which are not necessary to be canonical (cf. Proposition 2.1 and Example 2.2). When we take the generalized Cayley transforms $ and $' in the canonical fashion as in Example 2.2, that is, in the case when A = 1/2 and /i = 1, the following non-existence theorem has been proved by Ueno [7]. Theorem 2.1. f] Let A = 1/2,/x = 1 and m,m' > 2. Then there exists no proper harmonic map u : M —> M1 which extends to a C1 map up to the ideal boundary. However, if we choose A = /x = 1 and take the corresponding generalized Cayley transforms $ and $', then $ induces a new smooth structure on the half space model A^ x R+ of M, which is not diffeomorphic to the canonical one on N x {0}, and we have many proper harmonic maps from M to M'. Namely, the following existence theorem was proved by Donnelly [2]. Theorem 2.2. f2] Let A = /i = 1. Then, for any C1 map / between the ideal boundaries M(oo) = 5 2 m _ 1 and M'(oo) = S™''1 with nowhere vanishing energy density e(f), there exists a proper harmonic map u : M —> M' which assumes f continuously. 3. Asymptotic Behavior of Proper Harmonic Maps 3.1. Adapted
Frame
Fields
Let M = (G, g) be a fc-step Carnot space of dimension m and M' = (G' ,g') an Z-step Carnot space of dimension m', respectively. Recall that G and G' are simply connected, connected solvable Lie group that are semi-direct products G = N x R and G' = N' x R, where A' is a fc-step Carnot group and N' is an /-step Carnot group, and g and g' are left invariant metrics of strictly negative curvature, respectively. Also, by definition, the Lie algebras n and n' of N and N' are assumed to have graded Lie algebra
186
decompositions k
n = Y^ rii,
rij = {X 6 n | adiJ(X) = i\X},
i = 1 , . . . , k,
(3-D
T n
n
n
' = X^ i '
'i = (
I e n
' l adff'(X) = i/iX},
j = l,..., I,
3= 1
where the generators H and H' of the Lie algebra of the abelian group R are chosen as in Remark 2.1. Correspondingly, under generalized Cayley transforms <E> : N x .R+ -> G and $ ' : TV' x R+ ->• G', the pull-back metrics $*<7 and $'*' on the half space models TV x R+ and N' x R+ are expressed as 1 1 ^ 9= ^Xdn, + ^9n2
1 +••• + ^ 9 n
k
dy2 + ~r, (3.2)
5
y/2/i ffn; +
yl4lJ,
9n'2-\
I" ,2,,J 9n[ +
where y and y' denote the coordinates on their respective positive half lines R+, respectively. As in the case of complex hyperbolic spaces we looked at in Example 2.1, under the left translations of N, each subspace n* in the graded Lie algebra decomposition of n in (3.1) defines a distribution on N, that is, a subbundle of the tangent bundle of N. Indeed, each n;, being identified with a subspace of the tangent space of N at the identity, defines by left translations of N a subspace (rii)p of the tangent space of N at each p € TV. We denote these distributions on N also by n*, 1 < i < k, which may be regarded, under the identification of N x {0} with M(oo) \ {oo} as in Proposition 2.2, as defining a geometric structure on the ideal boundary Af(oo) of M. Similarly, we obtain the distributions nj, 1 < j < I, on N', which also define a geometric structure on the ideal boundary M'(oo) of M'. Now, let u : M -» M' be a proper C°° map from M to M'. Suppose that u extends to a continuous map u : M —> M' from the geometric compactification M = M U M(oo) of M to the geometric compactification M' = M' U M'(oo) of M'. We are then interested in necessary conditions to be satisfied by the boundary value / = u|Af (oo) : M(oo) -)• M'(oo) of u at the ideal boundary M(oo), when u is a harmonic map in the interior M.
187
In order to carry out computations in terms of local expressions, we take the half space models N x R+ and N' x R+ of M and M', and identify, via the corresponding generalized Cayley transforms, the geometric compactifications M and M 7 with (N x [0, oo)) U {00} and (N' x [0,00)) U {00'}, respectively. Here 00 G M(oo) and 00' E M'(co) denote the points at infinity defined by R+ directions in N x R+ and N' x R+ (see Proposition 2.2), respectively. We then define adapted frame fields {e^} on N x R+ and {e'a} on N' x R+ in the following way: Let eo = d/dy and ej, = d/dy'. Set n^ = dimn^ and n'P = dimnp, where 1 < A < A; and 1 < P < I. For each r u , choose an orthonormal basis { e ^ } , ! < i < UA, relative to the left invariant metric gnA, and extend them to left invariant vector fields on N. Note that, since we have [ n ^ n g ] C XIA+B a s in (2.3), we may write the structure constants of N as nA+B
[eAi,eBi]=
^2 aAA^)re{A+B)r,
l
and all other bracket products are trivial. Similarly, for each n'P, choose an orthonormal basis {e'p }, 1 < a < n'p, relative to the left invariant metric g'n, , and extend them to left invariant vector fields on N', with respect to which the structure constants of N' may be written as n
P +Q
7=1
all other bracket products being trivial. With respect to these frame fields, we denote the differential du and the tension field T(U) of u as
du
m —1 m'— 1
e
r u
m' — 1
= E E < * ® «> ( ) = E
ei
a=0
respectively, where {e,*} denotes the dual coframe field of {ej}. Also, as usual we denote the components of the metric g by gij — g(ei,ej) and the inverse matrix of (gij) by (g*-7). In terms of these adapted frame fields, as a direct consequence of (3.2), we have the following local expression for the tension field T(U), which will be convenient to investigate the asymptotic behavior of proper harmonic maps between Carnot spaces. Proposition 3.1. Let u : M —> M' be a C2 map from a k-step Carnot space M to an l-step Carnot space M1, which maps N x R+ to N' x R+.
188 Then, with respect to adapted frame fields, the local expression r(u)a tension field r(u) of u is given by
of the
m —1
A
i«)° = E s"(ei •«?) + (i - A E
• n*)y<
=0 771 — 1
u
(y'ou)-^9 (u°)2 m —1 i=0
I
np
P-i
/3=1
(3.3)
k
771 — 1
9U(ei • «f°) + (l - A E ^ ' ^ ) r f
T(U)P° = E ?=0
,4-1 771 — 1
2=0 n
-P
Q
n
P-f-Q k(*'+Q)-r„,Q/3„.( p +Q)-,
j=0
Q=l
where 1 < P < I and I < a
0=1
7=1
It should be remarked that in (3.3) the fourth term of T(u)Pa does not appear when P = I. 3.2. The case of X = fi To clear up the nature of the problem, let us first look at the case where A = /i. Namely, we suppose that on the half space models N x R+ and N' x R+ the pull-back metrics $*g and $'*g' in (3.2) blow up with the same growth orders at the boundaries N x {0} and N' x {0}, respectively. Also, for the sake of simplicity of argument, we set A = \i = 1. Let u : M —> M' be a proper C°° map from a fc-step Carnot space M to an /-step Carnot space M'. Suppose that u maps N x R+ to N' x R+ and extends to be a C 1 map up to the boundaries TV" x {0} and N' x {0}, that is, we now suppose that ueC°°(N
xR+,N'
x J ? + ) n C 1 ( 7 V x [0,oo),N' x [0,oo)).
Then we investigate to what extent the growth orders of the metrics $* and $'*' affect the asymptotic behavior of the tension field T(U) of u at the ideal boundaries.
189 First, with respect to adapted frame fields, we look at the e'0 component T°(U) of r(u). For each integer 1 < B < I, multiply the formula for T°(U) in (3.3) by (y' o u)2l~1y~2B and let y —> 0. Then we obtain Lemma 3.1. Let u : M —> M' be a proper C°° map from a k-step Carnot space M to an I-step Carnot space M', which maps N x R+ to N' x R+. Suppose that u extends to a Cx map up to the boundaries N X {0} and iV'x{0}. Then, for each 1 < B < l, with respect to adapted frame fields, the local expression r(u)° of the tension field T(U) of u satisfies the following. (1) As y —>• 0, the first three terms of T(U)° X (y1 ou)2l~ly~2B converges to 0 if B < I, and to
~{J2AnA)(u°0)21
ifB = l.
A=l
(2) The fourth term of T(U)° X (y' ou)2l~1y~2B
P=l-B+1
/3=1 (
m(B,P)
+ E P=l-B+1
where m(B,P)
is given by
nA
rip
E ^£E(^-V°«) , - p «2) 2 +°(i). A=l
i = l /3=1
= min{B + P -
l,k}.
Since r(u)° = 0 if u is a harmonic map, from Lemma 3.1 with B — 1, we can easily deduce the following condition satisfied by the Cl boundary value of a proper harmonic map. Lemma 3.2. Under the assumption of Lemma 3.1, ifu is a harmonic map, then it satisfies the following conditions on the boundary N x {0}. (1) If 1 = 1, then
( E ^ ) K ) 2 = EEK lp ) 2 A=l
i=0 0=1
(2) If I > 1, then for each 1 < i < n\ and 1 < /3 < n\ ulg = 0,
uf. = 0.
For instance, if u is a harmonic self-map of the complex hyperbolic space CHm, then, since k = I = 2 and n'2 = 1, Lemma 3.2 (2) states that u2 = 0
190 for 1 < i < n\ provided y = 0. Namely, in terms of the half space model N x R+ of CHm, the C1 boundary value / of u satisfies dfP({ni)P) C ( n i ) / ( p ) j 2 1 1
p£Nx{0},
2m_1
which means that / : S " " -+ 5 is a contact transformation on the ideal boundary 5 2 m _ 1 (cf. Example 2.1 and [2]). It should be remarked that in general Lemma 3.2 (2) implies Corollary 3.1. Let I > 2 and u : M -+ M' be a proper harmonic map from a k-step Carnot space M to an I-step Carnot space M', which maps N x R+ to N' x R+. Suppose that u extends to a Cl map up to the boundaries N x {0} and N1 x {0}. Then the boundary value f of u satisfies i-i
dfP((m)p) C XlK'W) for any p e J V x {0}. If a given proper harmonic map u has higher differentiability up to the boundaries, then we have further control of the asymptotic behavior of u at the boundary. More precisely, we obtain the following necessary conditions involving higher order derivatives of u at the boundary, which can be deduced from Lemma 3.1 with an induction argument. Proposition 3.2. Let I > 2 and 1 < r < I - 1. Let u : M -> M1 be a proper harmonic map from a k-step Carnot space M to an I-step Carnot space M', which maps N x R+ to N1 x i ? + . Suppose that u extends to a Cr map up to the boundaries N x {0} and N' x {0}. Then the following hold on N x {0}. (1) For each l-r + l
e0c.((y'o«)'-p^)=0,
0
(2) For each I - r + 1 < P
e?-((z/'°w/~ P u2) = 0 ,
0
Corollary 3.2. Under the assumption of Proposition 3.2, if in particular UQ 7^ 0 on the boundary N x {0} ; then the following hold on N x {0}. (1) For each l-r + l
0<s
+ l
P
e 0 • u /. = 0,
+ r-l. min{P -l + r, k},
0 < s < P + r - A - I.
191
Regarding the derivative UQ of u at the boundary, we can also deduce from Lemma 3.1 the following Lemma 3.3. Let u : M —> M' be a proper harmonic map from a k-step Carnot space M to an I-step Carnot space M', which maps N x R+ to N' x R+. If u extends to a Cl map up to the boundaries N x {0} and N1 x {0}, then it satisfies on N x {0}
( £ AnA) K)2/ - J2 p i > -1)!}-2 (4-1 • ((,' o uy-pu?))2 A=\
P=l I
min{P,A;}
0=1
nA
n'P
i=l
0=1
- E E pEY:{(l-Ayr2{ei0-A-((y'oUy-pu2)) P=\
A=l
=o.
In the case where k > I > 2, we can easily see from Lemma 3.3 that if
EE( i=l
u
'f)
2
^
0
on7Vx{0},
(3.4)
0=1
or equivalently the boundary value / of u satisfies dfP((ni)P)tJ2(n'i)f(p)>
P£Nx{0},
(3.5)
then UQ 7^ 0 at the boundary JV x {0}, since min{P, k} = P. Combining this with Corollary 3.2 with r = I — 1, we then obtain the following result. Proposition 3.3. Let k > I > 2. Let u : M —> M' be a proper harmonic map from a k-step Carnot space M to an I-step Carnot space M', which maps N x R+ to N' x R+. Suppose that u extends to a Cl map up to the boundaries N x {0} and N' x {0}, and satisfies (3.4) or equivalently (3.5). Then the following hold on N x {0}. (1) For each
2
UQ0
= 0,
(2) For each 2
0 < s < P - 2, 1 < 0 < n'P.
and 1 < A < min{P - 1, k}, 0<s
Now, we turn to other components T{U)PCX of the tension field T(U) to obtain further necessary conditions to be satisfied by the boundary value of a proper harmonic map u : M -> M'. To this end, in the remainder of
192
the argument, we suppose that u extends to a Cl map up to the boundaries N x {0} and N' x {0}, and satisfies
t6-0)
i=l 0=1
u°0^0
ifk
on N x. {0}. Namely, it is assumed in both cases that ujj ^ 0 on the boundary N x {0}. Under these assumptions, we multiply the formula for T(U)P" in (3.3) by (y' o u)2l~p~ly~2l+1, and let y —> 0. Then, invoking Corollary 3.2 and Proposition 3.3, we can deduce Lemma 3.4. Let u : M —> M' be a proper C°° map from a k-step Carnot space M to an l-step Carnot space M', which maps N x R+ to N' x R+. Suppose that u extends to a Cl map up to the boundaries N x {0} and N1 x {0}, and satisfies (3.6) on N x {0}. Then, as y —> 0, the local expression T(U) a of the tension field T(U) of u satisfies the following. (1) For 1 < P < I, the first three terms ofT(u)Pa x (y' ou)2l~p~ly~2l+l converges to
~{(P ~ I)!}"1 ( E AnA + p) (4)2l-P~1
(e^1 • uP«).
A=l
(2) For 1 < P < I - 1, the fourth term of converges to
T{U)P«
X
(y1
oU)2l'p~1y-2l+1
Q=l
0=1 7=1 l-P
m\n{Q,k)
+E
ci{P,Q,A)(ul)2l-p-2Q-1
E
Q=l HA
A=l n
Q
n
P+Q
x E E E i>Z£h^~A •«??)(ep+Q~A • « r o ) l ) . i=l /3=1
where Cl(P,Q,A)
7=1
= {(Q-A)l(P
+
Q-Ay.}-1.
193
As an immediate consequence of this Lemma, we then obtain the following conditions satisfied by the Cl boundary value of a proper harmonic map, since T(u)Pa = 0 if u is harmonic. Corollary 3.3. Under the assumption of Lemma 3.4, if u is a harmonic map, then it satisfies on the boundary N x {0}, 4 " 1 • u{,° = 0,
l
and
{(p -1)!}- 1 ( E
An
* + p) K) 2 ( " P ) (tf-1 • tf-) n
Z-P Q=l
Q "P+«
0 = 1 7=1
x[c1(P,g,l)(e?-1.^)(e^-1.ur^) min{Q,k} UA
= 0 A=l
t=l
/or 1 < P < / - 1 and 1 < a < n'p. To proceed further, we now exploit the integrability condition ddu = 0, which implies for any C2 map u that m—1
ej • u7 =
ei
• u] - E
m' — 1 e
ep(t i> e,-])uj[ -
p=0
E
< ([e'a> e fl]) u °"f •
a,0=O
From this identity we obtain for instance Pa
P-l nQ n'p- „ . . „,P« _ V ^ V " Y ^ ^ ?; <5/5 ? ,(^-0)-. Qf>(P-Qhuo Ai Q=l/3=1 7=1
Taking differentiation by eo successively and invoking Corollary 3.2 with r = I — 1, we then deduce Lemma 3.5. Let u : M —> M ' 6e a proper harmonic map from a k-step Carnot space M to an I-step Carnot space M', which maps N x JL|_ to N' x R+. Suppose that u extends to a Cl map up to the boundaries N x {0} and N' x {0}, and satisfies (3.6) on N x {0}. Then, for each 2 < P < I
194
and 1 < A < min{P — l,k}, the following holds on N x {0}. n
p_A
n
Q
P-Q
Z^ Zs Zs "QfsiP-Q)-, Q=l /3=1 7 = 1
x c2(P,Q,A)(e^1 where c2{P,Q,A)
P-A-V - .p _ ^ _
• u^)(%-A-Q
• uATQh
..
Relabel the indices of the identity in Lemma 3.5 and substitute it back into the identity in Corollary 3.3. Then we multiply the resulting equation by ep~x • uPa and sum it up over a, 1 < a < n'P, to yield
A=l l-P
n'p
n'Q
a=l
n'p+Q
Q = l a = l /3=1 7 = 1
l-P
min{Q,k}
P+Q-A
rip
n'Q n'p+Q
n'R
n'p+Q_R
nA
+E E E EEEE E E Q=l
A=l
R=l
a=l/3=l 7=1
/J=1
"=1
i=l
cUP,Q,A)c 2 (P + Q ) i ? ^ ) & ^ & ^ Q _ f i ) ^ K ) 2 ( ( - p - C ? ) x ( e r 1 • tf-) ( e ? " 1 • &)(<#-* = 0.
• uQ/M+Q'A~R
•
«$J+°-JR)-)
Observing that the summands of the second term are anti-symmetric in P and Q, it is not hard to see that the sum of the second term over P, 1 < P < Z — l , i s 0 . Similarly, since d(P, S - P, A)c2(5, R, A) = (P - l)\(S - A)-1
c3(S,P,A)c3(S,R,A),
where c3(S,P,A)
=
{(P-l)\(S-P-A)\}-\
it is also verified that the sum of the third term over P, 1 < P < I - 1,
195
yields I min{S-l,fc} n's UA
C-D'E E S=2
A=l
EE^-^K) 2 "^
7=1 i = l
x{i:\(S>P>A)£Bf,6j:(s_p)fl(^-1.<0(eoS-p-A-tx2-P,/,)}: <*=1 0 = 1
P=l
Hence we obtain on iV x {0}
Ei^-Dir^E^+^K) 2 ^!:^- 1 -^) 2 P=l ;
A=l min{S—1,*} n's n^
a=l
+ E E EE^rW"^ S=2
A=l
7=1 i = l
^bfLls_^(^.^-)(4-p-A-u%-p^)}i
x {^c3(S,P,A)± = 0.
Since all terms in the sum are nonnegative and u° ^ 0 by assumption, it follows that u satisfies o n J V x {0} ep~l
-up«=0
for each 1 < P < I — 1 and 1 < a < n'p, and
Q=l
a=l
0=1
for each 2 < P < Z , l < y l < min{P - 1, A;}, where 1 < 7 < rip, 1 < i < n A Noting that c2(P,Q,A) = (P - A - l)\c3(P,Q,A) and substituting back into Lemma 3.5 then yields ep~A
• up:
= 0
for 2 < P < I and 1 < A < min{P — 1,/c}. Summing up all these results, we obtain the following necessary conditions for the boundary value of a proper harmonic map between Carnot spaces. Theorem 3.1. Let u : M —>• M' be a proper harmonic map from a kstep Carnot space M to an I-step Carnot space M', which maps N x R+ to N' x R+. Suppose that u extends to a Cl map up to the boundaries N x {0}
196
and N' x {0}, and satisfies (3.6) on N x {0}. Then the following hold on N x {0}. (1) For each 1 < P < I, er0 • UQ" = 0,
0 < r < P - 1, 1 < a < n'P.
(2) For each 2