Interlacing of zeros of weakly holomorphic modular forms

By Paul Jenkins and Kyle Pratt

Abstract

We prove that the zeros of a family of extremal modular forms interlace, settling a question of Nozaki. Additionally, we show that the zeros of almost all forms in a basis for the space of weakly holomorphic modular forms of weight for interlace on most of the lower boundary of the fundamental domain.

1. Introduction and main results

A natural question in studying functions of a complex variable is to determine the location of the zeros of a function; an especially interesting case occurs when the locations of the zeros follow a strong pattern. Many modular forms have zeros satisfying such properties. The most well-known such result comes from F. Rankin and Swinnerton-Dyer Reference 8, who proved that all zeros of the classical Eisenstein series in the standard fundamental domain lie on the circular arc on the lower boundary of .

When two functions have zeros that lie on the same arc, we say that the zeros of two functions interlace if every zero of one function is contained in an open interval whose endpoints are zeros of the other function, and each such interval contains exactly one zero. Gekeler conjectured that the Eisenstein series satisfies such interlacing properties in Reference 3, and Nozaki Reference 7 proved that the zeros of interlace with the zeros of by improving the bounds used by Rankin and Swinnerton-Dyer. For the modular function given by the action of the th Hecke operator on , Jermann Reference 6 extended work of Asai, Kaneko, and Ninomiya Reference 1 to prove that the zeros of interlace with the zeros of . In this paper, we prove interlacing for a family of holomorphic modular forms for , all of whose zeros in lie on the arc .

Denote by the space of holomorphic modular forms of weight , and write for the larger space of weakly holomorphic modular forms (i.e. poles are allowed at the cusps) of weight . For such weights , we write , with and . Duke and the first author Reference 2 introduced a canonical basis for whose elements are defined by

where, as usual, . They approximated on the boundary arc by a trigonometric function to prove the following theorem locating the zeros of these basis elements.

Theorem 1.1 (Reference 2, Theorem 1).

If , then all the zeros of in lie on .

The condition in Theorem 1.1 that is not sharp, but for all large enough weights , the zeros of at least one of the do not all lie on . The theorem is generally false when is a cusp form; one example is the form .

Getz studied a subset of these basis elements in Reference 4. We call this family of modular forms “gap functions”, as they are the holomorphic modular forms with the maximum possible gap in their -expansions. For any even weight , these functions are defined as

These gap functions have application to the theory of extremal lattices and questions in coding theory; see for example Reference 5.

Our first result proves that the zeros of these gap functions, all of which lie on , interlace.

Theorem 1.2.

Let be an even integer with . Then the zeros of interlace on with the zeros of on .

This theorem settles a question of Nozaki, who suggested that the zeros of these functions “interlace in some range” (see Reference 7, Example 1.1).

Additionally, we are able to partially extend our results to the larger class of functions . The result is the following theorem.

Theorem 1.3.

Let and fix (resp. ). Then the zeros of interlace with the zeros of (resp. ) on the arc

for (resp. ) large enough.

As the methods used in the proof of Theorem 1.2 do not entirely apply to the case in which is nonzero, they do not give interlacing on all of , although preliminary computations suggest that such interlacing generally holds. We leave this as an open problem.

2. Background

We begin by defining some notation. The standard fundamental domain for is

Let , and for even integers let be the usual weight Eisenstein series

where is the th Bernoulli number and . We let

be the usual modular function in and be the weight 12 cusp form

With this notation, the basis elements can be explicitly constructed as

where is a polynomial in of degree such that has the correct Fourier expansion.

The proof of Theorem 1.1 in Reference 2 depends on integrating a generating function for the to obtain

where and is a counterclockwise circle in the -plane centered at 0 with sufficiently small radius. We fix for some in the interval and change variables . Then for some we have

We move the contour downward to a height of . As we do so, each pole in the region

contributes to the value of the integral; these poles occur when is equivalent to under the action of . If a pole occurs with real part , we modify the contour to include small semicircles in the usual way.

We allow to vary depending on . This is to ensure that if is close to , then the residue term from does not appear. We choose if , picking up residues from and , and we let if , picking up an additional residue at . The overlap of the two intervals is necessary to obtain interlacing of the zeros. Applying the Residue Theorem and taking absolute values as in the proof of Theorem 1.1 in Reference 2, we obtain

when and

when .

Note that for any modular form of weight , the function is real-valued for , so the left-hand sides of (Equation 1) and (Equation 2) are absolute values of real-valued functions of . Thus, these inequalities give approximations for the modular forms by the trigonometric function , whose zeros are proved to interlace in the next section. To prove Theorem 1.2, we show that the right-hand sides of equations (Equation 1) and (Equation 2) are exponentially decaying functions in the weight , preserving the interlacing for sufficiently large. To prove interlacing in the first interval is straightforward. On the other hand, the additional residue term in the second interval shifts the zeros of away from the zeros of the cosine function, necessitating more care. Computing the interlacing of the zeros of the for all smaller proves interlacing for all .

The proof of Theorem 1.3 proceeds along similar lines, using the fact that the right-hand sides of (Equation 1) and (Equation 2) are also exponentially decaying in .

3. Interlacing for cosine functions

In this section, we show that the cosine functions obtained in the residue calculation have zeros that interlace. We define

As we deal with variants of in which either is replaced by or is replaced by , we similarly define

and

When it is clear from context, we write to mean either or . Note that has one more zero in than does .

We first prove that the zeros of and interlace on .

Lemma 3.1.

If , then the zeros of interlace on with the zeros of and with the zeros of .

Proof.

It is clear that in order to have interlacing, the following four conditions are sufficient:

The first zero in belongs to .

The last zero in belongs to .

The zeros of and in are never equal.

Between two consecutive zeros of there is exactly one zero of .

We prove each of these assertions in turn. It is helpful to work with , because

so and have the same zeros. At the endpoints of , we have

and

By taking derivatives, we see that and are monotonically increasing on for and that for all . Thus, equations (Equation 6) and (Equation 7) imply that for all , we have

Since , the first zeros of and on occur when and are equal to , where is the first odd integer greater than . Let be the first zero of on so that . Then by (Equation 8) and (Equation 9), so the first zero of occurs before the first zero of . This proves the first of the assertions.

The proof of the second assertion follows similarly, though we must make adjustments depending on . To see that the zeros cannot be equal, we set ; equality can hold only for not in .

There can be at most one zero of between any two consecutive zeros of ; otherwise, must increase faster than between the zeros, a contradiction. Let be two consecutive zeros of so that for some integer . Applying (Equation 9) shows that , so there must exist a point in the interval with and . A similar argument shows that a zero of appears between every two zeros of in .

In showing interlacing for the , we will need bounds on the distances between the zeros of the approximating cosine functions. The following proposition gives a preliminary estimate on the distances between zeros.

Proposition 3.2.

Suppose . The distance between two consecutive zeros of is less than or equal to .

Proof.

The result follows from a simple application of the mean value theorem. We bound from beneath on by . Note that the condition on means that this derivative is always positive. As moves from one zero of to the next, must increase by ; the lower bound on the derivative gives an upper bound on the distance between zeros.

We use Proposition 3.2 to prove a stronger result.

Lemma 3.3.

Suppose that , and if , suppose that . The shortest distance in between a zero of and a zero of is between the first zero of and the first zero of or between the last zero of and the last zero of .

The lemma is clearly true when has only one zero in and is an immediate consequence of the following proposition.

Proposition 3.4.

Suppose that . Let be three consecutive zeros of in and be two consecutive zeros of in such that . If , suppose that . Then and .

This proposition says that as we examine an increasing sequence of intervals whose endpoints are zeros of , the zero of in each interval moves farther from the left-hand side of the interval and closer to the right-hand side of the interval.

Proof.

We begin with the case where we increase by 12 by comparing the derivatives and . If , these derivatives are constant and for any . Since increases by on and on and increases by on , we conclude that and , giving the desired inequalities.

If , then and are both decreasing functions of , as goes from to on . If for some , then has decreased by at least 6 on the interval . Since is bounded below by on , this can happen only if .

Since , by Proposition 3.2 it is clear that

for satisfying . Thus, we must have , and since are decreasing, it must be true that for all . Since increases by on and increases by on , we conclude that and .

To prove that , we show that . We have

If , this is less than , so for all . Noting that and increase by on the appropriate intervals as before, we conclude that .

We now handle the case in which we increase by 1. We begin by noting that

and Proposition 3.2 gives an upper bound on of

so the most that could decrease between and is

Now , and

if , which is always true when . When this condition on holds, for all in the interval , and both and increase by on the appropriate intervals, so we get the desired inequalities.

4. Proof of Theorem 1.2

Recall that Theorem 1.2 is a statement about , so we use Lemmas 3.1 and 3.3, when needed, with . We proceed with two cases, depending on the value of .

Suppose that . From Equation 1 it is clear that may be approximated by if we convert the right-hand side into a suitably decreasing function of . Lemma 3.1 shows that the zeros of interlace, so the zeros of will also interlace if is sufficiently close to .

Consider the quotient of functions on the right-hand side of Equation 1. By Proposition 2.2 of Reference 4, we know that for , we have

and we check computationally that for , we have

Together, for and in the appropriate ranges, these give

Duke and the first author proved that for ,

Taking Equation 1, Equation 10, and Equation 11 together shows that

Using the relation , it is clear that , so we define . We see that .

We now compute how far the zeros of can stray from the zeros of . Suppose that

for some constant , and let satisfy Then a zero of appears in the interval , where To get an upper bound on , consider the line intersecting through the points The concavity of near implies that this line lies between and the -axis. Therefore, if is a point at which the value of the line is , then . The absolute value of the slope of the line is , and it follows that .

Lemma 3.3 gives us a lower bound on the distances between the zeros of and ; this lower bound is when and when . This can be seen by considering what happens at the endpoints of for mod 4. For instance, if , then at we have that is an integer multiple of and needs to increase only by before has a zero in . When , we see that must increase by before a zero occurs.

Replacing with , we solve the inequality

which is true when . This means that when , the zeros of differ from the zeros of by an amount which is less than half the minimum distance between zeros of and . The zeros of and therefore lie in disjoint, interlacing intervals, and must interlace on for .

Now let . The method of the previous case must be modified, as we are dealing with a different approximating function for , given by

We require the following lemma.

Lemma 4.1.

The zeros of interlace with the zeros of on .

Proof.

Lemmas 4.1, 4.2, 4.3, 4.4, and 4.5 in Reference 7 prove that the zeros of interlace with the zeros of ; adding does not change the order of the zeros. The function is closely related to Nozaki’s function , defined as

In using Nozaki’s lemmas we may simply take .

The term is monotonically increasing, is very small for smaller , and tends rapidly to 1 for close to . This residue term shifts the zeros of away from the zeros of , but for large the effect is negligible unless is very near .

As before, we need a lower bound on the distance between the zeros of and . An easily adapted lemma from Nozaki (Reference 7, Lemma 4.1) shows that for a zero of and the corresponding zero of , we have

if . This fact follows from the observations that and for . We will use this fact frequently to estimate quantities involving zeros of .

Let denote a zero of and an adjacent zero of . There are two cases to consider: intervals of type , and intervals of type . We will obtain lower bounds on the length of intervals of both types.

Consider first the intervals. We may view these essentially as intervals defined by zeros of and , along with some zero shifts due to the presence of the term. By Proposition 3.4, the shortest such interval is the first after .

We proceed by cases, according to the congruence class of . Suppose that , so that . We solve to find the smallest such that is a zero of and similarly find the next zero of . If is even, then the distance between these cosine zeros is

while if is odd, then the interval has length

Therefore, by Proposition 3.4 and Equation 13, a lower bound on the length of a interval is given by

which is positive for all .

This argument works for each value of . We have the following lower bounds for intervals of type:

The method for handling the intervals cannot be easily adapted for intervals, so we use a different approach. Our general strategy of proof involves the function . If , we show that this function is monotonically increasing or decreasing on the interval . We then obtain a lower bound on , which gives the change in the value of this function over the interval, and use the trivial bound on the derivative of given by

to find a lower bound for . On the other hand, if , we may simply use as a lower bound.

To see that is monotonic on the interval when , we note that the interval is contained in the interval . On this larger interval, the absolute value of the derivative of ranges from to . The absolute value of the derivative of , on the other hand, is bounded above by its value at an upper bound for the largest possible of ; this value is at most . Thus, the derivative of the cosine term dominates in the interval, and is monotonic; it follows that is monotonic on .

We now bound when . There are three cases to consider, based on the value of mod 12, since the behavior of depends heavily on . In each case there are two subcases, since for this type of interval the zeros of and either both shift to the left or both shift to the right from the zeros of and .

Case : .

Consider first the subcase in which and are increasing at consecutive zeros, so that the zeros are shifted to the left by adding the extra term. For any with a lower bound on is given by , since is positive for all . By trigonometric identities this is equal to .

The function has zeros on at , where is a nonnegative integer. In this case, with the zeros shifted left, is even. On the other hand, the zeros of are at , with . Given this restriction on , it is straightforward to confirm that

implying that the zeros and of and both lie between the same two zeros of .

When the zeros of and are shifted left, the zero of is greater than . Because of the parity of , the derivative of is negative at , implying that the function is positive on all of and that a lower bound on is given by

which simplifies to

With the condition , we find that

and since is even this implies . Therefore, a lower bound on when are shifted left is given by , since is increasing in .

Now suppose that are shifted right. Then a lower bound on is given by for some . Ignoring for the moment the exponential terms, we find a lower bound on by arguing as above to replace by . Inserting this for in the function and simplifying as before, we have

Because is now odd, we find that , so this is bounded from beneath by .

Now we consider the term . Since is increasing on and is decreasing on , an upper bound on the absolute value of is given by

Since is decreasing in for fixed , a lower bound on is given by

These two lower bounds, along with Equation 15, give a lower bound on for of the smaller of and

It is not difficult to verify the positivity of . Comparing with the appropriate quantities from Equation 14, we find that is indeed a lower bound on the zero distance.

We saw earlier that near zeros of we have . If it is true that , then this gives an upper bound of on the distance a zero of can travel. Performing calculations as in the case where , we may take to be . We then solve the inequality

which holds for .

Case : .

This and the following case follow very similarly to the one above. Again we find that lie between two zeros of , and we handle separately the cases in which and are shifted left or right. If we define

then a lower bound on zero distance is given by

Here is positive for , and comparison with Equation 14 shows that is a lower bound on zero distance over the range of consideration. Similarly as above, we bound from beneath the absolute value of the derivative of , and find it is bounded by . We find that the inequality

holds for .

Case : .

Proceeding as above, a lower bound on the distance between zeros of and is given by

where

The inequality here is

which holds for .

Comparing our results from the two intervals, we find that the zeros of interlace with the zeros of on the lower boundary of the fundamental domain for . We have confirmed computationally that the zeros interlace for , and we have an appropriate intersection between our two intervals, since Proposition 3.2 implies that this intersection contains at least two zeros when . It follows that the zeros of and interlace on the lower boundary of the fundamental domain.

5. Proof of Theorem 1.3

For convenience we restate Theorem 1.3.

Theorem 1.3.

Let and fix (resp. ). Then the zeros of interlace with the zeros of (resp. ) on the arc

for (resp. ) large enough.

Our proof follows the outlines of the proof above; the most significant differences involve the lower bounds on the distances between zeros. We take linear approximations to and and use those approximations to derive lower bounds on the distance between zeros. We require the hypotheses for Lemma 3.3 to hold; we then need only find such a bound near and .

We first determine the bound for zeros near . Taking the first order Taylor series approximation for gives us

When we increase by 12, the linear approximations to and have the same error term.

Write . Note that is increasing and positive on , since is concave down. Let be the first zeros of and in , respectively, and let be the first zeros on of and . We then have, for integers and ,

Now we find the slopes of the lines between and and between and , and apply the Mean Value Theorem. The slope can be taken to be the value of the derivative at a point in the interval, and by the proof of Proposition 3.4 the derivative of is greater than the derivative of in the appropriate intervals for large enough. Thus, for large we have that

which implies . This in turn implies that , so the distance between the zeros of and is less than the distance between the zeros of and . Computing the distance between the zeros of and , we get a lower bound of

for the distance between zeros near .

The argument for increasing by 1 is not exactly analogous because the error terms are no longer identical. We use the Taylor series approximation for and use the fact that near we have close to its first order approximation.

Assume since the case is similar. Additionally, assume . If we find that the lower bound on the zeros is greater than the lower bound when . Bounding from beneath the derivative of , we see that because , the first zero of in is less than . By Taylor’s Theorem,

for some . This gives the inequality

for large enough with respect to .

Since the first zero of in is at , we see that the first zero of in is less than . Thus, a lower bound on the distance between the first zeros of and is given by

which is positive for large enough with respect to .

We use similar arguments for the lower bound near and find that the lower bound between zeros, whether we increase or increase , is given by a decreasing rational function in and .

Now we pick , which is fixed for the remainder of the proof. From equations Equation 1 and Equation 2 and the proof of Theorem 1.1 we obtain

where . We do so by comparing the bounds for the two different intervals and choosing our bound to be larger than both of them. Note that each term of the right side is exponentially decaying in both and .

Suppose more generally that . We want to derive an upper bound in terms of on the distance a zero of can be from a zero of . We will do this by bounding from beneath the absolute value of the derivative of on intervals around each of its zeros. If is small enough, the zeros of must lie in these intervals, and we may argue as in the proof of Theorem 1.2 to obtain a bound involving an exponentially decaying quantity.

Suppose that for some . We calculate that

Using the above bounds, we must bound from below on a suitable interval. Trivially bounding the derivative of , we see that is an insufficient change in the absolute value for to reach an extreme. Thus we consider the interval . Standard formulas give

where and . When or is large enough, is very near 1, so and . When is fixed and increases, we see that approaches , so is close to . When is fixed and increases, approaches and is close to or greater than , since may range from to . With this bound at the endpoints of our interval, we let be a positive constant smaller than such that for all under consideration with fixed and increasing (or fixed and increasing) and , we have

Letting be the minimum of the four lower bounds on the distance between zeros we calculated above and replacing with our exponential quantities from before, we find that the zeros interlace when

Because the left-hand side has exponential decay in and while the right-hand side is a rational function of and , we see that the inequality holds for all but finitely many and , so the zeros of interlace with the zeros of or on for sufficiently large.

Mathematical Fragments

Theorem 1.1 (Reference 2, Theorem 1).

If , then all the zeros of in lie on .

Theorem 1.2.

Let be an even integer with . Then the zeros of interlace on with the zeros of on .

Theorem 1.3.

Let and fix (resp. ). Then the zeros of interlace with the zeros of (resp. ) on the arc

for (resp. ) large enough.

Equation (1)
Equation (2)
Lemma 3.1.

If , then the zeros of interlace on with the zeros of and with the zeros of .

Equation (6)
Equation (7)
Equations (8), (9)
Proposition 3.2.

Suppose . The distance between two consecutive zeros of is less than or equal to .

Lemma 3.3.

Suppose that , and if , suppose that . The shortest distance in between a zero of and a zero of is between the first zero of and the first zero of or between the last zero of and the last zero of .

Proposition 3.4.

Suppose that . Let be three consecutive zeros of in and be two consecutive zeros of in such that . If , suppose that . Then and .

Equation (10)
Equation (11)
Equation (13)
Equation (14)
Equation (15)

References

Reference [1]
Tetsuya Asai, Masanobu Kaneko, and Hirohito Ninomiya, Zeros of certain modular functions and an application, Comment. Math. Univ. St. Paul. 46 (1997), no. 1, 93–101. MR1448475 (98e:11052),
Show rawAMSref \bib{AKN97}{article}{ author={Asai, Tetsuya}, author={Kaneko, Masanobu}, author={Ninomiya, Hirohito}, title={Zeros of certain modular functions and an application}, journal={Comment. Math. Univ. St. Paul.}, volume={46}, date={1997}, number={1}, pages={93--101}, issn={0010-258X}, review={\MR {1448475 (98e:11052)}}, }
Reference [2]
W. Duke and Paul Jenkins, On the zeros and coefficients of certain weakly holomorphic modular forms, Pure Appl. Math. Q. 4 (2008), no. 4, 1327–1340. Special issue in honor of Jean-Pierre Serre, DOI 10.4310/PAMQ.2008.v4.n4.a15. MR2441704 (2010a:11068),
Show rawAMSref \bib{DJ08}{article}{ author={Duke, W.}, author={Jenkins, Paul}, title={On the zeros and coefficients of certain weakly holomorphic modular forms}, journal={Pure Appl. Math. Q.}, volume={4}, date={2008}, number={4}, pages={1327--1340. Special issue in honor of Jean-Pierre Serre}, issn={1558-8599}, review={\MR {2441704 (2010a:11068)}}, doi={10.4310/PAMQ.2008.v4.n4.a15}, }
Reference [3]
Ernst-Ulrich Gekeler, Some observations on the arithmetic of Eisenstein series for the modular group , Festschrift: Erich Lamprecht, Arch. Math. (Basel) 77 (2001), no. 1, 5–21, DOI 10.1007/PL00000465. MR1845671 (2002f:11050),
Show rawAMSref \bib{Gek01}{article}{ author={Gekeler, Ernst-Ulrich}, title={Some observations on the arithmetic of Eisenstein series for the modular group ${\rm SL}(2,{\mathbb {Z}})$}, journal={Festschrift: Erich Lamprecht, Arch. Math. (Basel)}, volume={77}, date={2001}, number={1}, pages={5--21}, issn={0003-889X}, review={\MR {1845671 (2002f:11050)}}, doi={10.1007/PL00000465}, }
Reference [4]
Jayce Getz, A generalization of a theorem of Rankin and Swinnerton-Dyer on zeros of modular forms, Proc. Amer. Math. Soc. 132 (2004), no. 8, 2221–2231, DOI 10.1090/S0002-9939-04-07478-7. MR2052397 (2005e:11047),
Show rawAMSref \bib{Getz04}{article}{ author={Getz, Jayce}, title={A generalization of a theorem of Rankin and Swinnerton-Dyer on zeros of modular forms}, journal={Proc. Amer. Math. Soc.}, volume={132}, date={2004}, number={8}, pages={2221--2231}, issn={0002-9939}, review={\MR {2052397 (2005e:11047)}}, doi={10.1090/S0002-9939-04-07478-7}, }
Reference [5]
Paul Jenkins and Jeremy Rouse, Bounds for coefficients of cusp forms and extremal lattices, Bull. Lond. Math. Soc. 43 (2011), no. 5, 927–938, DOI 10.1112/blms/bdr030. MR2854563,
Show rawAMSref \bib{JR11}{article}{ author={Jenkins, Paul}, author={Rouse, Jeremy}, title={Bounds for coefficients of cusp forms and extremal lattices}, journal={Bull. Lond. Math. Soc.}, volume={43}, date={2011}, number={5}, pages={927--938}, issn={0024-6093}, review={\MR {2854563}}, doi={10.1112/blms/bdr030}, }
Reference [6]
Jonas Jermann, Interlacing property of the zeros of , Proc. Amer. Math. Soc. 140 (2012), no. 10, 3385–3396, DOI 10.1090/S0002-9939-2012-11212-2. MR2929008,
Show rawAMSref \bib{Jer12}{article}{ author={Jermann, Jonas}, title={Interlacing property of the zeros of $j_n(\tau )$}, journal={Proc. Amer. Math. Soc.}, volume={140}, date={2012}, number={10}, pages={3385--3396}, issn={0002-9939}, review={\MR {2929008}}, doi={10.1090/S0002-9939-2012-11212-2}, }
Reference [7]
Hiroshi Nozaki, A separation property of the zeros of Eisenstein series for , Bull. Lond. Math. Soc. 40 (2008), no. 1, 26–36, DOI 10.1112/blms/bdm117. MR2409175 (2009d:11070),
Show rawAMSref \bib{Noz08}{article}{ author={Nozaki, Hiroshi}, title={A separation property of the zeros of Eisenstein series for ${\rm SL}(2,\mathbb {Z})$}, journal={Bull. Lond. Math. Soc.}, volume={40}, date={2008}, number={1}, pages={26--36}, issn={0024-6093}, review={\MR {2409175 (2009d:11070)}}, doi={10.1112/blms/bdm117}, }
Reference [8]
F. K. C. Rankin and H. P. F. Swinnerton-Dyer, On the zeros of Eisenstein series, Bull. London Math. Soc. 2 (1970), 169–170. MR0260674 (41 #5298),
Show rawAMSref \bib{RSD70}{article}{ author={Rankin, F. K. C.}, author={Swinnerton-Dyer, H. P. F.}, title={On the zeros of Eisenstein series}, journal={Bull. London Math. Soc.}, volume={2}, date={1970}, pages={169--170}, issn={0024-6093}, review={\MR {0260674 (41 \#5298)}}, }

Article Information

MSC 2010
Primary: 11F11 (Holomorphic modular forms of integral weight), 11F03 (Modular and automorphic functions)
Author Information
Paul Jenkins
Department of Mathematics, Brigham Young University, Provo, Utah 84602
jenkins@math.byu.edu
MathSciNet
Kyle Pratt
Department of Mathematics, Brigham Young University, Provo, Utah 84602
kvpratt@gmail.com
MathSciNet
Communicated by
Ken Ono
Journal Information
Proceedings of the American Mathematical Society, Series B, Volume 1, Issue 7, ISSN 2330-1511, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2014 by the authors under Creative Commons Attribution-Noncommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S2330-1511-2014-00010-9
  • MathSciNet Review: 3211795
  • Show rawAMSref \bib{3211795}{article}{ author={Jenkins, Paul}, author={Pratt, Kyle}, title={Interlacing of zeros of weakly holomorphic modular forms}, journal={Proc. Amer. Math. Soc. Ser. B}, volume={1}, number={7}, date={2014}, pages={63-77}, issn={2330-1511}, review={3211795}, doi={10.1090/S2330-1511-2014-00010-9}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.