Coefficient estimates on weighted Bergman spaces

John E. McCarthy ∗ Washington University, St. Louis, Missouri 63130, U.S.A.

0 Introduction

Let A denote normalized area measure for the unit disk D in C. The Bergman 2 2 La is the sub-space of the L (A) consisting of functions n 2 that are also analytic in D. The monomials z are orthogonal in La, and √ 1 P∞ n 2 have n+1 ; so a f(z) = n=0 anz is in La if and P∞ 2 1 PN n only if n=0 |an| n+1 < ∞, and if this is so, the partial sums n=0 anz are converging to f. The weighted Bergman spaces normally studied are obtained by replacing the measure dA(z) by the radial measure dAα(z) = (1 − |z|2)αdA(z); in these spaces the monomials are again orthogonal, and a function is approximable in norm by the partial sums of its power series at zero. In this paper we are interested in studying non-radial weights of the 2 ∞ form |m(z)| dAα(z), where m is the modulus of a function in H , the space of bounded analytic functions on D. There are two different ways of generalizing the Bergman space to these 2 2 weights. We shall use La(|m| Aα) to denote the space of analytic functions 2 2 2 2 on D that also lie in L (|m| Aα), and P (|m| Aα) to denote the closure of the ∗The author was partially supported by the National Science Foundation grant DMS 9296099.

1 2 2 polynomials in L (|m| Aα). These two spaces are, in general, different. If m 2 2 2 has no zeroes, L (|m| Aα) is just the set of quotients {f/m : f ∈ La}; but this 2 2 2 2 coincides with P (|m| Aα) only when 1/m is in P (|m| Aα), which in turn is 2 equivalent to requiring that 1 lie in the La closure of {mp : p a }, i.e. that m be a cyclic vector for multiplication by z (this is discussed in Section 3 below). When m has an infinite number of zeroes, we do not know 2 2 2 2 when P (|m| Aα) is all of La(|m| Aα). 2 2 2 2 We are interested in obtaining, for functions in P (|m| Aα) and La(|m| Aα), estimates on the size of the Taylor coefficients, and on the rate of growth as the boundary is approached. Our principal results are the following.

∞ 2 2 Theorems 2.1, 2.6, 1.4 Let m be in H . Let f be in La(|m| Aα). Then

√ |fˆ(n)| = eO( n log n).

2 2 If f is in La(|m| A), then

√ |fˆ(n)| = eO( n).

If f is in P 2(|m|2A), then

√ |fˆ(n)| = eo( n).

Theorem 3.6 Assume m in H∞ is zero-free, and suppose the Taylor series 1 1 P∞ n expansion of m is given by m(z) = n=0 αnz . Then

v u n ˆ 2 2 uX 2 sup{|f(n)| : f ∈ La(|m| A), kfk ≤ 1} = t (n + 1 − j)|αj| j=0

2 1 P 2(|m|2A)

2 2 2 Let PLa denote the orthogonal projection from L (A) onto La, and for m 2 2 ∞ La 2 2 La in H define the co-analytic Toeplitz operator Tm¯ : La → La by Tm¯ f =

2 PLa (mf ¯ ). If σ is Lebesgue measure on the circle, the classical 2 2 2 2 H is the closure of the polynomials in L (σ) (so H = P (σ)), and if PH2 denotes projection from L2(σ) onto H2, then define the co-analytic Toeplitz H2 2 2 H2 operator Tm¯ : H → H by Tm¯ f = PH2 (mf ¯ ). Just as in the Hardy space case [2, 5], the function f is in the range of the 2 La Toeplitz operator Tm¯ if and only if there exists some constant C such that, for all polynomials p, | R pfdA¯ |2 ≤ C R |p|2|m|2dA, i.e. if the functional D p 7→ R pfdA¯ is bounded on P 2(|m|2A). If this functional is bounded, we D shall say f is in the dual of P 2(|m|2A). The important point is that if P∞ ˆ n 2 2 P∞ n f(z) = n=0 f(n)z is in the dual of P (|m| A), and g(z) = n=0 gˆ(n)z is in P 2(|m|2A), then the sum

∞ 1 X gˆ(n)fˆ(n) n=0 n + 1 must converge; so knowing a function f is in the dual of P 2(|m|2A) means that all functions g satisfy n + 1 gˆ(n) = o( ). fˆ(n)

The common range of co-analytic Toeplitz operators on H2 was charac- terized in [5] in terms of the Fourier coefficients fˆ(n) of the function f:

Theorem 1.1 The function f is in the range of every co-analytic Toeplitz operator on the Hardy space if and only if there exists a constant c > 0 such √ that fˆ(n) = O(e−c n).

In the expository article [7] I stated that Theorem 1.1 was also true in the Bergman space. However, whereas the proof of the necessity of the decay

3 condition on the Taylor coefficients is valid, the dismissal of the proof of sufficiency seems unjustified. Using a result of Lotto and Sarason [4], we can prove the theorem. Lotto and Sarason’s result is that the set Y , of analytic √ functions f satisfying fˆ(n) = O(e−c n) for some c, is what they call Toeplitz stable, i.e. it is a subspace with the properties that ∞ H2 (i) If f is in Y and m is an outer function in H , then Tm¯ f is in Y . ∞ H2 (ii) If f is in Y and m is an outer function in H , then f = Tm¯ h for some h in Y .

P∞ n 2 Theorem 1.2 Let f(z) = n=0 anz be in La. A necessary and sufficient condition that f lie in the range of every non-zero co-analytic Toeplitz oper- √ 2 −c n ator on La is that, for some c > 0, an = O(e ).

Proof: Necessity: We use the following relation between the Toeplitz op- 2 2 La 2 H 2 erator Tm¯ on La and the operator Tm¯ on H : If ∞ ∞ H2 X n X n Tm¯ cnz = dnz n=0 n=0 then ∞ ∞ 2 La X n X n Tm¯ cn(n + 1)z = dn(n + 1)z , (1.3) n=0 n=0 P∞ 2 whenever n=0 |cn| (n + 1) < ∞. Now for each m there is some function 2 P∞ n P∞ 2 1 La g(z) = n=0 bnz , where n=0 |bn| n+1 < ∞, with Tm¯ g = f. By (1.3), H2 P∞ 1 n P∞ 1 n P∞ 1 n Tm¯ n=0 bn n+1 z = n=0 an n+1 z . So n=0 an n+1 z is in the range of every 1 co-analytic operator on the Hardy space, and so by Theorem 1.1 {an }, √ n+1 −c n and hence also {an}, is O(e ). ∞ P∞ an n Sufficiency: Fix m in H . Let f1(z) = n=0 n+1 z . Then f1 is in Y , P∞ n H2 so there is some function g1(z) = n=0 bnz in Y with Tm¯ g1 = f1. Let 2 P∞ n 2 La g(z) = n=0 bn(n + 1)z ; then g is in La and by (1.3) Tm¯ g = f. 2

Corollary 1.4 Let m be in H∞ and g be in P 2(|m|2A). Then √ |gˆ(n)| = eo( n).

4 2 2 2 La(|m| Aα)

2 2 We first estimate the growth of the Taylor coefficients of functions in La(|m| Aα). We shall use C to denote a generic constant, that may change from one line to the next. Instead of requiring m to be in H∞ we shall allow it to range over the larger class A−α consisting of all functions g that are holomorphic iθ C on D and satisfy |g(re | ≤ (1−r)α .

−α 2 2 Theorem 2.1 Let m be in A , and f be in La(|m| Aα). Then

iθ O( 1 log( 1 )) |f(re )| = e 1−r 1−r , 0 < r < 1 (2.2)

and √ |fˆ(n)| = eO( n log n), n > 1. (2.3)

1+r 3+r Proof: Let ρ be between 2 and 4 , and let

it iθ it ρe + re P iθ (ρe ) = < re ρeit − reiθ be the Poisson kernel. Then

Z 2π iθ 1 it it 1 log |f(re )| ≤ Preiθ (ρe )[log |fm(ρe )| + log ]dt. 2π 0 |m(ρeit)|

Without loss of generality, m(0) 6= 0 (otherwise, this zero can be factored out). Then by Jensen’s formula

Z 2π 1 log− |m(ρeit)|dt ≤ C + α log 0 1 − ρ so

Z 2π iθ 4 1 1 it it log |f(re )| ≤ [C + α log ] + Preiθ (ρe ) log |fm(ρe )|dt, 1 − r 1 − r 2π 0

5 and hence

Z 2π iθ C 4α log 1 it it |f(re )| ≤ e 1−r e 1−r 1−r Preiθ (ρe ) log |fm(ρe )|dt. (2.4) 0 Multiply both sides of (2.4) by (1 − ρ2)α and integrate with respect to ρ 1+r 3+r from 2 to 4 . This gives

iθ C C 4α log 1 |f(re )| ≤ e 1−r e 1−r 1−r , (2.5) (1 − r)α+1

which yields (2.2). Cauchy’s formula tells us that for any r < 1,

ˆ 1 iθ |f(n)| ≤ n sup |f(re )|. r 0≤θ≤2π

√1 Choosing r = 1 − n in (2.2) yields (2.3). 2

Note that if α = 0 the bound in (2.5) is sharper, which enables us to eliminate the log-term in the coefficient estimate:

∞ 2 2 Corollary 2.6 If m is in H and f is in La(|m| A), then √ |fˆ(n)| = eO( n).

Let F + be the space of functions f satisfying

√ |fˆ(n)| = eo( n). (2.7)

This space was first studied by Yanagihara [13]; he showed that a function f is in the algebra F + if and only if

( o(1) ) |f(z)| = e 1−|z| . (2.8)

6 Yanagihara also showed that the Smirnov class is contained in F + [14] (see also [6]), so in particular it contains the reciprocal of every outer function. It also contains, by a result of J. Shapiro and A. Shields [10], those singular inner functions whose associated measures are non-atomic; we include a proof for completeness. (For definitions and basic properties of the Smirnov class, inner and outer functions see e.g. the book [1].) Proposition 2.9 Suppose the H∞ function m is the product of an outer function, h, and a singular inner function, u, whose associated measure is 1 + non-atomic. Then m is in F . Proof: Assume |m| is given by Z iθ − log |u(z)| = Pz(e )dµ(θ). (2.10) T Define l(s) = sup{µ(I): Ian interval, σ(I) = s}, and notice that lims→0 l(s) = 0 precisely when µ is non-atomic. Without loss of generality, take z = r to be positive, and break the integral (2.10) into the integrals over the sets where 1 (i) |θ| < (1 − r) 4 , and 1 (ii) |θ| ≥ (1 − r) 4 . 1 1+r 4 The first integral is less than {l[2(1 − r) ]}{ 1−r }, and the second is less 1 1 i(1−r) 4 i(1−r) 4 √ 1 than kµkPr(e ). For r close to 1, Pr(e ) is approximately 1−r . 1 Therefore m satisfies (2.8), as required. 2 2 2 We want to consider conditions on m which imply that La(|m| A) is contained in F +. Theorem 2.11 Let m in H∞ be the product of an outer function, a singular inner function whose associated measure is non-atomic, and a Blaschke prod- ∞ P∞ − 1 uct whose zero-set {wn}n=1 satisfies n=0(1 − |wn|)(log (1 − |wn|)) 2 < ∞. 2 2 Let f be in La(|m| A). Then √ |fˆ(n)| = eo( n).

7 Proof: Let B be the Blaschke product with the same zero-set as m, so m = Bm1 where m1 is the product of a non-atomic singular inner function 2 2 + and an outer function. Let g = fm1 ∈ La(|B| A). If we can show g is in F , then so is f, because 1 is in F + by the preceding remarks. To show g is in m1 F +, it is enough to show that Z (log− |B|)2dA < ∞, (2.12) D for this implies that Z (log+ |g|)2dA < ∞, D and this in turn implies that g is in F + by a theorem of Stoll [12].

So, suppose B is a Blaschke product with zero-set {wn}. Then

Z 1 1 2 R P∞ 1−w¯nz 2 [ | log |B(z)|| dA(z)] 2 = [ ( log | |) dA(z)] 2 D n=0 z−wn D 1 P∞ R 1−w¯nz 2 ≤ [ (log | |) dA(z)] 2 (2.13) n=0 D z−wn

Now let us estimate R (log | 1−w¯nz |)2dA(z). D z−wn 1 As each term in (2.13) is finite, we can assume each |wn| ≥ 2 . For convenience, assume w is positive, and make the change of variables ζ = iθ z−w re = 1−wz . Then

Z Z 2 2 1 − wz 2 1 1 2 (1 − w ) (log | |) dA(z) = (log ) iθ 4 rdrdθ D z − w π D r |1 − wre | Z 1 1 1 + (rw)2 = 2(1 − w2)2 (log )2 rdr(2.14) 0 r (1 − (rw)2)3

1 Break the integral (2.14) into two pieces: from 0 to e , where the integrand 1 is bounded by some constant C1 independent of w, and from e to 1. For the 1 latter integral, use the inequality log r ≤ 2(1 − wr). One gets that (2.14) is bounded by C(1 − w)2(1 + log−(1 − w)), so by hypothesis 2.13 is finite, as required. 2

8 3 Kernel Functions

Let us fix some function m in H∞. Every function in P 2(|m|2A) is analytic on D, and evaluation at each point of D is a bounded linear functional. Therefore m for each w in D, there exists some kernel function kw satisfying Z m 2 f(w) = f(z)kw (z)|m| dA(z). D m m As |f(w)| ≤ kfkP 2(|m|2A)kkw kP 2(|m|2A), we want to estimate kkw kP 2(|m|2A) = q m kw (w). When m ≡ 1, the kernel function, which we shall then call kw, is 1 known to be kw(z) = (1−wz¯ )2 . Therefore for any polynomial p, Z Z m p(w) = p(z)kw(z)dA(z) = p(z)kw (z)m(z)m(z)dA(z). (3.1) D D m 2 2 m 2 m As kw is in P (|m| A), mkw is in La, and (3.1) says that kw andmmk ¯ w have the same inner product with polynomials, i.e.

2 1 m La m = k = P 2 mmk¯ = T mk (3.2) (1 − wz¯ )2 w La w m¯ w

m Thus to find kw , we must solve (3.2). Consider the function 1 1 1 Km(z) = , (3.3) w m(w) m(z) (1 − wz¯ )2

2 2 m and let us suppose for the time being that it lies in P (|m| A); then Kw m must in fact be kw , for

Z Z 1 1 1 2 1 1 p(z) 2 |m(z)| dA(z) = p(z)m(z) 2 dA(z) D m(w) m(z) (1 − wz¯ ) m(w) D (1 − wz¯ ) = p(w).

m 2 2 1 2 2 When is Kw in P (|m| A)? This occurs if and only if m is in P (|m| A), which in turn happens if and only if m is a cyclic vector for the Bergman 2 ∞ shift (the operator of multiplication by z on La). For an H function m,

9 a necessary and sufficient condition that it be cyclic for the Bergman shift is that it not vanish on D and that the singular measure associated with its singular inner factor put no mass on any BCH set (necessity was proved by H. Shapiro [9], sufficiency independently by B. Korenblum [3] and J. Roberts [8]; see the survey article by A. Shields [11]). A BCH set (which stands for Beurling-Carleson-Hayman set) is a compact subset of the circle, of Lebesgue P∞ measure zero, for which n=0 σ(In) log(σ(In)) > −∞, where {In} are the disjoint open arcs of the complement. m Thus, for m a cyclic vector for the Bergman shift, (3.3) gives us kw explicitly, and hence the norm of evaluation at w. If m is not cyclic, then m m 2 2 2 2 kw is the projection of Kw from L (|m| A) onto P (|m| A); unfortunately, we do not know how to calculate this explicitly. Summarising the above, we have proved:

Theorem 3.4 Let m be in H∞, and suppose m(w) 6= 0. Then the reproduc- ing kernel on P 2(|m|2A) for evaluation at w is given by

m m kw (z) = PP 2(|m|2A)[Kw (z)] 1 1 1 = P 2 2 [ ]. P (|m| A) m(w) m(z) (1 − wz¯ )2

m 2 2 If m is cyclic for the Bergman shift then Kw lies in P (|m| A), and the norm 1 1 of evaluation at the point w is |m(w)| 1−|w|2 ; otherwise the norm is strictly smaller than this.

2 2 A similar discussion applies to La(|m| A). The reproducing kernel at w m m 2 2 2 2 is PLa(|m| A)Kw , and Kw lies in La(|m| A) if and only if m is zero-free. m If m vanishes at w, it is easy to work out what Kw should be. A calcu- lation gives:

Scholium 3.5 Suppose m has a zero of order n at w. Then

n m (n + 1)! 1 z k (z) = P 2 2 . w P (|m| A) m¯ (n)(w) m(z) (1 − wz¯ )n+2

10 m Once one knows the reproducing kernels kw for evaluation at w, one also knows the kernels that, when integrated against a polynomial, give the values of some derivative of the polynomial; for if Z m 2 p(w) = p(z)kw (z)|m(z)| dA(z), then Z (n) (n) m 2 Dw p(w) = p(z)Dw kw (z)|m(z)| dA(z),

(n) where Dw denotes differentiating n times with respect to w. This observa- tion enables us to calculate explicitly the growth rate of the Taylor coefficients of a function in P 2(|m|2A) when m is cyclic for the Bergman shift:

Theorem 3.6 Assume m in H∞ is cyclic for the Bergman shift, and suppose 1 1 P∞ n the Taylor series expansion of m is given by m(z) = n=0 αnz . Then q ˆ 2 2 2 2 2 sup{|f(n)| : f ∈ P (|m| A), kfk ≤ 1} = |αn| + 2|αn−1| + ... + (n + 1)|α0| (3.7)

m 2 2 Proof: Let kw,n be the function in P (|m| A) that gives the value of the nth derivative at w: Z (n) m 2 p (w) = p(z)kw,n(z)|m(z)| dA(z).

m th m As remarked above, kw,n is just the n derivative, with respect tow ¯, of kw , so from Theorem 3.4 we have 1 1 1 km (z) = D(n) w,n w¯ m(z) m(w) (1 − wz¯ )2 ! 1 n (n) 1 1 = [ Dw¯ ( )( ) m(z) 0 m(w) (1 − wz¯ )2 ! n (n−1) 1 1 + Dw¯ ( )Dw¯( ) + ... 1 m(w) (1 − wz¯ )2 ! n 1 (n) 1 + ( )(Dw¯ )] n m(w) (1 − wz¯ )2

11 In particular, when w = 0, we get n ! m 1 X n (n−j) 1 j k0,n(z) = [ Dw¯ ( )|w¯=0(j + 1)!z ] m(z) j=0 j m(w) n 1 X j = [n! (j + 1)¯αn−jz ]. m(z) j=0

1 m The left-hand side of (3.7) is n! kk0,nk. Calculating:

m 2 n m kk0,nk = Dz (k0,n(z))|z=0 n n 1 X j = n!Dz [ (j + 1)¯αn−jz ]|z=0 m(z) j=0 n ! X n (n−j) 1 = n! [Dz ( )(j + 1)¯αn−jj!]|z=0 j=0 j m(z) n 2 X 2 = (n!) (j + 1)|αn−j| . j=0 Taking the square-root, and dividing by n!, one gets that the norm of the functional that sends a function to its nth Taylor coefficient at 0 is the right- hand side of (3.7), as desired. 2 The same technique as above enables one to estimate the growth rate of Taylor coefficients of functions in H2(|m|2σ). One gets:

1 P∞ n Theorem 3.8 Let m be an outer function, and suppose m(z) = n=0 αnz . Then n ˆ 2 2 X 2 1 sup{|f(n)| : f ∈ H (|m| σ), kfk ≤ 1} = [ |αj| ] 2 . j=0

Note that this is a qualitative improvement on the “asymptotic Szeg¨o theorem,” Theorem 3.1 in [6], which just states that, for any c > 0, the √ c n left-hand side is less than or equal to C1e .

Acknowledgement: The author would like to thank the referee for the key ideas in the proof of Theorem 2.1, which allowed the theorem to be significantly strengthened from its original form.

12 References

[1] John B. Garnett. Bounded Analytic Functions. Academic Press, New York, 1981.

[2] H. Helson. Large analytic functions II. In Cora Sadosky, editor, Analysis and partial differential equations. Marcel Dekker, Basel, 1990.

[3] B. I. Korenblum. Cyclic elements in some spaces of analytic functions. Bull. Amer. Math. Soc., 5:317–318, 1981.

[4] B.A. Lotto and D. Sarason. Multipliers of de Branges-Rovnyak spaces. Indiana Math. J., 42(3):907–920, 1993.

[5] J.E. McCarthy. Common range of co-analytic Toeplitz operators. J. Amer. Math. Soc., 3(4):793–799, 1990.

[6] J.E. McCarthy. Topologies on the Smirnov class. J. Funct. Anal., 104(1):229–241, 1992.

[7] J.E. McCarthy. Weighted Bergman spaces. In K. Jarosz, editor, Func- tion spaces, pages 301–306. Marcel Dekker, 1992.

[8] J. Roberts. Cyclic inner functions in the Bergman space and weak outer functions in Hp, 0 < p < 1. Illinois Math. J., 29(1):25–38, 1985.

[9] H.S. Shapiro. Some remarks on weighted polynomial approximation of holomorphic functions. Mat. Sbornik, 73(115):320–330, 1967.

[10] J.H. Shapiro and A.L. Shields. Unusual topological properties of the Nevanlinna class. Amer. J. Math., 97(4):915–936, 1973.

[11] A.L. Shields. Cyclic vectors in spaces of analytic functions. In S.C. Power, editor, Operators and function theory, pages 315–349. D. Reidel, 1985.

13 [12] M. Stoll. Mean growth and Taylor coefficients of some topological alge- bras of analytic functions. Ann. Polon. Math., XXXV:139–158, 1977.

[13] N. Yanagihara. The containing Fr´echet space for the class N +. Duke Math. J., 40:93–103, 1973.

[14] N. Yanagihara. Mean growth and taylor coefficients of some classes of functions. Ann. Polon. Math., XXX:37–48, 1974.

14