Zeros of a one-parameter family of harmonic trinomials

By Michael Brilleslyper, Jennifer Brooks, Michael Dorff, Russell Howell, and Lisbeth Schaubroeck

Abstract

It is well known that complex harmonic polynomials of degree may have more than zeros. In this paper, we examine a one-parameter family of harmonic trinomials and determine how the number of zeros depends on the parameter. Our proof heavily utilizes the Argument Principle for Harmonic Functions and involves finding the winding numbers about the origin for a family of hypocycloids.

1. Introduction

Let be a continuous complex-valued harmonic function, where both and are real-valued harmonic functions. Such a function can be expressed as , where and are analytic.

A familiar fact about analytic functions is that they are sense-preserving (or orientation-preserving) at all points at which the derivative is not zero. Complex-valued harmonic functions, however, may be sense-preserving in one region and sense-reversing in another. Lewy’s Theorem Reference 13 implies that is locally univalent and sense-preserving if and only if and the dilatation function of , defined by , satisfies . The dilatation function can be viewed as a measure of how far is from being analytic; for analytic functions, and is the well-studied class of quasiconformal functions Reference 11.

Much work has been done examining the similarities and differences between analytic and harmonic functions. Many familiar results for analytic functions hold for complex-valued harmonic functions with only slight modifications. The harmonic analog of the Argument Principle for Analytic Functions is a case in point Reference 7.

Argument Principle for Harmonic Functions: Let be a Jordan domain with boundary . Suppose that is harmonic on , continuous on , on , and there are no zeros of in for which . Then the total change in the argument of as is traversed in the positive direction is , where is the number of zeros of in counted according to their multiplicity.

The multiplicity of a zero of a complex-valued harmonic function is defined via its power series expansion about the zero. That is, let

where and . If is in the sense-preserving region of the plane, then and the order of the zero is . If is in the sense-reversing region of the plane, then and the order of the zero is defined to be . For more information about complex-valued harmonic functions, see Reference 4, Reference 5, and Reference 6.

A tantalizing problem is to determine the number of zeros of a complex-valued harmonic polynomial. For analytic polynomials, the Fundamental Theorem of Algebra stipulates that the number of zeros counting multiplicity equals the degree of the polynomial. In the harmonic case, the results are not so straightforward. If is a harmonic polynomial for which the degree of is , the degree of is , and , then Sheil-Small Reference 17 conjectured that the maximum number of zeros of is . Wilmshurst Reference 18 proved this conjecture and provided an example to show that is sharp. Independently, Peretz and Schmid Reference 16 also proved the conjecture, while Bshouty et al. Reference 3 provided another example to establish the sharpness of the bound. Wilmshurst further conjectured that in the more restrictive case in which , has at most zeros. Khavinson and Swiatek Reference 9 looked at the subclass of harmonic polynomials related to gravitational lensing. They showed that if the degree of is , then the number of zeros is bounded by , which satisfies Wilmshurst’s conjecture. However, Lee et al. Reference 12 later showed that Wilmshurst’s conjecture is not true in general.

Recently, Melman Reference 14 investigated trinomials of the form , where , , , and . He gleaned information relating to the location of the zeros of . In a similar fashion, Brilleslyper and Schaubroeck Reference 1, Reference 2, considered the family of trinomials , where , , and derived a formula for the number of zeros of located on the unit circle. Howell and Kyle Reference 8 then proved a conjecture in Reference 2 to determine the number of zeros of this same trinomial in the interior and exterior of the unit circle. In this paper, we look at a harmonic equivalent of these trinomials. That is, we consider the family of harmonic trinomials

where , , , and . We are interested in how the number of zeros of changes as varies. Our main theorem is the following.

Theorem 1.1.

Let be as in Equation 1.2 and let . There exist critical values , with , such that

(a)

if , then has distinct zeros,

(b)

if for some , then has distinct zeros, and

(c)

if , then has distinct zeros.

We will prove this theorem in the following section. In the process we also show that and . The following example and accompanying figure provide an illustration.

Example 1.2.

Consider , so that, with the notation of Theorem 1.1, , , and . As Figure 1 illustrates, when , the trinomial has zeros. When , corresponding to the case , has zeros. Finally, when , the number of zeros has increased to .

Theorem 1.1 states that the behavior seen for is typical: the number of zeros of increases from to as increases. The -values at which the number of zeros increases will be called the critical -values.

The curve in the plane that separates the sense-preserving and sense-reversing regions for plays a significant role in our analysis. A straightforward computation gives that if and only if , with being sense-reversing on the interior of this circle and sense-preserving on its exterior. Because of its central role, we make the following definition.

Definition 1.3.

The critical circle is

Note that in Figure 1(A) there are 5 zeros outside of , in Figure 1(B) there are 6 zeros outside of and 1 zero inside, and in Figure 1(C) there are 8 zeros outside of and 3 zeros inside. Recall from the Argument Principle for Harmonic Functions that a zero in a sense-preserving region has a positive order while a zero in a sense-reversing region has a negative order. So, in each case in Figure 1, the sum of the orders of the zeros is 5, which is the value of for . In this paper we prove that transitions between numbers of zeros occur when passes through the origin. Furthermore, we show that, for , the trinomial has no zeros inside , whereas for , it has zeros inside .

2. Proof of Theorem 1.1

To prove Theorem 1.1 we make use of a series of lemmas. First, keeping in mind that the zeros of may include zeros of both positive and negative order, Lemma 2.1 concludes that, for all , the sum of the orders of the zeros of equals . Lemma 2.2 then provides a short proof that every zero of has order or . It is also necessary to find the winding number of the image of the critical circle about the origin, because this information relates to the number of zeros in the sense-reversing region. Towards that end, Lemma 2.4 shows that the image is a certain type of hypocycloid that is centered at and whose size is affected only by . Further, the geometry of the hypocycloid reveals that the number of zeros can only change with additional zeros being balanced inside and outside the critical circle . Finally, Lemma 2.5 shows that has distinct intersections with the real axis to the right of its center at , and hence counts the winding number of around the origin for the various values of . The proof of our theorem follows by combining these lemmas.

Lemma 2.1.

Let be as in Equation 1.2. For sufficiently large, the winding number of the image of under around the origin is . Thus, for all , the sum of the orders of the zeros of is .

Proof.

The proof follows immediately from a standard Rouché-type argument by comparing with for sufficiently large .

We next demonstrate that the order of each zero is either or . Consequently, counting the number of distinct zeros of is equivalent to counting the number of zeros according to multiplicity.

Lemma 2.2.

All zeros of have order or .

Proof.

The trinomial can be written as . Let be a zero of . We know that since . The series expansions of and about are finite series since and are polynomials, where in accordance with the notation of equation Equation 1.1, and , neither of which is zero. Thus the order of the zero at is either if or if .

As already noted, the image of the critical circle plays an important role in our analysis. We show that is a hypocycloid of type per the following definition.

Definition 2.3 (Reference 10).

A hypocycloid centered at the origin is the curve traced by a fixed point on a circle of radius rolling inside a larger origin-centered circle of radius . The curve is given by the parametric equations

If the ratio is written in reduced form as , then the hypocycloid has cusps, and each arc connects cusps that are away from each other in a counterclockwise direction. Such a hypocycloid is called a hypocycloid, and the range of values to trace the entire hypocycloid is .

Lemma 2.4.

The image of the critical circle is an hypocycloid centered at . The value of affects only the size of the hypocycloid.

Proof.

Evaluating and splitting into real and imaginary parts gives

and

We make the substitutions and . As a consequence, we have , , and . Comparing with Definition 2.3, we see that the equations for and describe a hypocycloid centered at instead of the origin. We observe that the ratio does not depend upon the constant , and the entire hypocycloid is traced for .

Because the smaller circle has more than half the radius of the larger circle. Thus, although the inner circle travels in a counterclockwise direction, the hypocycloid traces around its center in a clockwise direction.

It follows from Lemma 2.4 that in Example 1.2 with , is an hypocycloid. Figure 2 exhibits this hypocycloid for three different values of .

Note that any two cusps that are cusps apart in the counterclockwise direction can also be viewed as being cusps apart in the clockwise direction. For the hypocycloids of Figure 2, each cusp is connected by an arc to cusps that are 3 cusps away, one in the counterclockwise and one in the clockwise direction.

In the following argument we make use of the geometry of the hypocycloid. For any value of the number of zeros of in the sense-reversing region of the plane can be determined by finding the winding number of the hypocycloid about the origin. For sufficiently small, the hypocycloid is entirely to the left of the origin, so . As increases, remains constant until the hypocycloid passes through the origin, which occurs for the first time when equals the first critical value . As continues to increase, there will be other critical values for which the hypocycloid passes through the origin. These are the only places where the winding number can change. Counting the number of critical values thus requires us to first count the number of intersections of an hypocycloid with the portion of the real axis to the right of its center.

Lemma 2.5.

The hypocycloid has distinct intersections with the real axis to the right of its center. These intersections, in turn, correspond to the critical -values with , at which intersects the origin. The winding number, , of around the origin is as follows:

(a)

If , then .

(b)

If for , then .

(c)

If , then .

Proof.

We first show that the number of distinct intersections of the hypocycloid with the real axis to the right of its center is . The rightmost cusp of the hypocycloid corresponds to in the parameterization . Hence, solving the equation for  gives the first transition value. We easily obtain . Thus for , the hypocycloid lies entirely to the left of the origin.

We label the rightmost cusp of the hypocycloid with , and the other cusps in a counterclockwise direction with . Note that labels on cusps must be computed modulo . For example, the cusp with label can also be referred to using the label . Two arcs emanate from each cusp, and each arc connects cusps that are apart. Since cusp  is in the upper half plane, and connects to cusp on the real axis. Based on the labeling of cusps and the geometry of the hypocycloid, each cusp labeled has an arc that connects to a cusp in the lower half plane and crosses the real axis to the right of the center of the hypocycloid. No other arcs can cross this portion of the real axis. When combined with the arc from cusp  to cusp  we see that each cusp labeled is the end point of one arc that intersects the real axis to the right of center.

The hypocycloid is symmetric with respect to the real axis, however, so not all intersections with the real axis are unique. Let denote the arc of the hypocycloid that connects cusps and . Then for each cusp with the arcs and pass through the same point on the real axis. These two arcs are distinct except in the case where . This modular equation is equivalent to which is only satisfied when is even and . We also observe that an arc which is its own mirror image must be the leftmost intersection of the hypocycloid with the real axis that lies to the right of the center.

Thus, for odd, every distinct intersection involves a pair of arcs, except for the arc . So in this case there are distinct intersections. For the case even, there are distinct intersections. These formulations can be combined using the floor function to give distinct intersections for all values of .

Now we compute the winding number , using the crossing condition as stated in Reference 15. By that condition, we only need to consider a path from the origin going to the right, and count how many curves we must cross to “escape” the hypocycloid. Each curve that we cross that is moving from our right to left as we travel away from the origin contributes to . Combining the crossing condition with the discussion above, we see that as increases, the winding number changes according to the pattern , which is expressed in part (b). Once is large enough that all intersections with the real axis to the right of the center of the hypocycloid are also to the right of the origin, we have . Note that in the case odd and even, the last change in the winding number is from to .

Each change in the winding number occurs at a critical value of . Thus there are such values.

We are now prepared to prove the main theorem.

Proof of Theorem 1.1.

By Lemma 2.1 and the Argument Principle for Harmonic Functions, the sum of the orders of the zeros of is always . Furthermore, by Lemma 2.2, all zeros of have order or . Since is completely to the left of zero when , we conclude that for such , has only zeros. We next discuss . Since , there is one zero inside , so there must be zeros outside of , for a total of zeros. Now, when for , we have . We apply the Argument Principle to say that the number of zeros in the sense-preserving region of the plane must be . Thus the total number of zeros is , as desired. Similarly, since for , we have that has zeros in this case.

Finally, we note that the upper and lower bounds for and can be found directly using the triangle inequality by examining for what values goes through the origin. The condition that implies that

Applying the triangle inequality, we have that

or, equivalently, that

This, in turn, implies that

Since all the transitions from zeros to zeros occurs within a small range of real numbers.

3. Areas for further investigation

(1)

In this paper, we showed how the number of zeros changes as varies for , but we did not prove anything about the location of the zeros. What can be shown about the location of the zeros for ?

(2)

What are the exact values for for ?

(3)

The subclass of harmonic polynomials is related to gravitational lensing, and it has been shown Reference 9 that if the degree of is , then the number of zeros is bounded by . Could the approach of this paper be used to improve that bound or to show that it is sharp?

(4)

Our proof of Theorem 1.1 relies heavily on the fact that the image of the critical circle is a hypocycloid with a cusp on the real axis. If is complex but not real, the image curve no longer has this simple geometry. What is the analogue of Theorem 1.1 for complex ?

(5)

What can be proven about the number of zeros of other families of harmonic polynomials?

Figures

Figure 1.

Zeros of for , and circles

Figure 1(a)

(A) Zeros of

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \begin{axis} [ width=1.9in, height=1.9in, axis x line=center, x axis line style=thick, axis y line=center, y axis line style=thick, xtick={-1,1}, ytick={-1,1}, tick style={thick, black}, tick label style={font=\tiny}, clip=false, enlargelimits=false, xmin=-1.8, xmax=1.8, ymin=-1.8, ymax=1.8, ] \node[anchor= west] at (axis cs:1.8,0){$x$}; \node[anchor= south] at (axis cs:0,1.8){$y$}; \node[anchor=west] at (axis cs:0.5,0.7){\scriptsize$|z|=\left(\frac{3}{5}\right)^{\frac{1}{2}}$}; \addplot[only marks, mark=*, mark size=1.75pt, blue] coordinates{ (0.83762, 0) (-1.0251, -0.505977) (-1.0251, 0.505977) (0.447145, -1.14267) (0.447145, 1.14267) }; \draw(0,0) circle [radius=sqrt(3/5)]; \end{axis} \end{tikzpicture}
Figure 1(b)

(B) Zeros of

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \begin{axis} [ width=1.9in, height=1.9in, axis x line=center, x axis line style=thick, axis y line=center, xtick={-1,1}, xticklabels={\hspace*{-1.5em} -1, \hspace*{0.5em} 1}, xticklabel shift={-3pt}, ytick={-1,1}, yticklabels={\raisebox{-10pt} {-1}, \raisebox{5pt} 1}, yticklabel style={anchor=east}, yticklabel shift=-3pt, tick style={thick, black}, tick label style={font=\tiny}, y axis line style=thick, clip=false, enlargelimits=false, xmin=-1.8, xmax=1.8, ymin=-1.8, ymax=1.8, ] \node[anchor= west] at (axis cs:1.8,0){$x$}; \node[anchor= south] at (axis cs:0,1.8){$y$}; \node[anchor=west] at (axis cs:0.55,0.9){\scriptsize$|z|=\left(\frac{4.5}{5}\right)^{\frac{1}{2}}$}; \addplot[only marks, mark=*, mark size=1.75pt, blue] coordinates{ (0.779881, 0) (-1.18401, -0.536298) (-1.18401, 0.536298) (0.508828, -1.26723) (0.508828, 1.26723) (1.06278, -0.366282) (1.06278, 0.366282) }; \draw(0,0) circle [radius=sqrt(4.5/5)]; \end{axis} \end{tikzpicture}
Figure 1(c)

(C) Zeros of

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \begin{axis} [ width=1.9in, height=1.9in, axis x line=center, x axis line style=thick, axis y line=center, xtick={-1,1}, ytick={-1,1}, tick style={thick, black}, tick label style={font=\tiny}, y axis line style=thick, clip=false, enlargelimits=false, xmin=-1.8, xmax=1.8, ymin=-1.8, ymax=1.8, ] \node[anchor= west] at (axis cs:1.8,0){$x$}; \node[anchor= south] at (axis cs:0,1.8){$y$}; \node[anchor=west] at (axis cs:0.7,1.3){\scriptsize$|z|=\left(\frac{9}{5}\right)^{\frac{1}{2}}$}; \addplot[only marks, mark=*, mark size=1.75pt, blue] coordinates{ (0.66251, 0) (-1.61461, -0.682156) (-1.61461, 0.682156) (-0.646221, -1.54462) (-0.646221, 1.54462) (-0.335166, -0.643315) (-0.335166, 0.643315) (0.676235, -1.64531) (0.676235, 1.64531) (1.58512, -0.642104) (1.58512, 0.642104) }; \draw(0,0) circle [radius=sqrt(9/5)]; \end{axis} \end{tikzpicture}
Figure 2.

The hypocycloids for , and .

Figure 2(a)

(A)

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \begin{axis} [ width=2in, height=2in, axis x line=center, x axis line style=thick, axis y line=center, y axis line style=thick, ytick={-1,1}, xtick={1}, extra x ticks={-1}, extra x tick labels={}, tick style={thick, black}, tick label style={font=\tiny}, clip=false, enlargelimits=false, xmin=-2, xmax=2, ymin=-2, ymax=2, ] \node[anchor= west] at (axis cs:2,0){$x$}; \node[anchor= south] at (axis cs:0,2){$y$}; \addplot[domain=0:2*pi, blue, smooth, samples=200, line width=2pt] ( { 9/25*sqrt(3/5)*cos(deg(5*x)) + 3/5*sqrt(3/5)*cos(deg(3*x)) - 1 }, { 9/25*sqrt(3/5)*sin(deg(5*x)) - 3/5*sqrt(3/5)*sin(deg(3*x)) } ); \end{axis} \end{tikzpicture}
Figure 2(b)

(B)

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \begin{axis} [ width=2in, height=2in, axis x line=center, x axis line style=thick, axis y line=center, y axis line style=thick, ytick={-2,2}, xtick={2}, extra x ticks={-1}, extra x tick labels={}, tick style={thick, black}, tick label style={font=\tiny}, clip=false, enlargelimits=false, xmin=-3, xmax=3, ymin=-3, ymax=3, ] \node[anchor= west] at (axis cs:3,0){$x$}; \node[anchor= south] at (axis cs:0,3){$y$}; \addplot[domain=0:2*pi, blue, samples=200, line width=2pt] ( { 9/4*sqrt(3/2)*(9/25*sqrt(3/5)*cos(deg(5*x)) + 3/5*sqrt(3/5)*cos(deg(3*x))) - 1 }, { 9/4*sqrt(3/2)*(9/25*sqrt(3/5)*sin(deg(5*x)) - 3/5*sqrt(3/5)*sin(deg(3*x))) } ); \end{axis} \end{tikzpicture}
Figure 2(c)

(C)

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \begin{axis} [ width=2in, height=2in, axis x line=center, x axis line style=thick, axis y line=center, y axis line style=thick, ytick={-12,12}, xtick={12}, extra x ticks={-1}, extra x tick labels={}, tick style={thick, black}, tick label style={font=\tiny}, clip=false, enlargelimits=false, xmin=-16, xmax=16, ymin=-16, ymax=16, ] \node[anchor= west] at (axis cs:16,0){$x$}; \node[anchor= south] at (axis cs:0,16){$y$}; \addplot[domain=0:2*pi, blue, samples=200, line width=2pt] ( { 81*cos(deg(3*x))/(5*sqrt(5)) + 243*cos(deg(5*x))/(25*sqrt(5)) - 1 }, { -648*(2+3*cos(deg(2*x)))*(sin(deg(x))^3)/(25*sqrt(5)) } ); \end{axis} \end{tikzpicture}

Mathematical Fragments

Equation (1.1)
Equation (1.2)
Theorem 1.1.

Let be as in Equation 1.2 and let . There exist critical values , with , such that

(a)

if , then has distinct zeros,

(b)

if for some , then has distinct zeros, and

(c)

if , then has distinct zeros.

Example 1.2.

Consider , so that, with the notation of Theorem 1.1, , , and . As Figure 1 illustrates, when , the trinomial has zeros. When , corresponding to the case , has zeros. Finally, when , the number of zeros has increased to .

Lemma 2.1.

Let be as in Equation 1.2. For sufficiently large, the winding number of the image of under around the origin is . Thus, for all , the sum of the orders of the zeros of is .

Lemma 2.2.

All zeros of have order or .

Definition 2.3 (Reference 10).

A hypocycloid centered at the origin is the curve traced by a fixed point on a circle of radius rolling inside a larger origin-centered circle of radius . The curve is given by the parametric equations

If the ratio is written in reduced form as , then the hypocycloid has cusps, and each arc connects cusps that are away from each other in a counterclockwise direction. Such a hypocycloid is called a hypocycloid, and the range of values to trace the entire hypocycloid is .

Lemma 2.4.

The image of the critical circle is an hypocycloid centered at . The value of affects only the size of the hypocycloid.

Lemma 2.5.

The hypocycloid has distinct intersections with the real axis to the right of its center. These intersections, in turn, correspond to the critical -values with , at which intersects the origin. The winding number, , of around the origin is as follows:

(a)

If , then .

(b)

If for , then .

(c)

If , then .

References

Reference [1]
Michael A. Brilleslyper and Lisbeth E. Schaubroeck, Locating unimodular roots, College Math. J. 45 (2014), no. 3, 162–168, DOI 10.4169/college.math.j.45.3.162. MR3207562,
Show rawAMSref \bib{bs1st}{article}{ author={Brilleslyper, Michael A.}, author={Schaubroeck, Lisbeth E.}, title={Locating unimodular roots}, journal={College Math. J.}, volume={45}, date={2014}, number={3}, pages={162--168}, issn={0746-8342}, review={\MR {3207562}}, doi={10.4169/college.math.j.45.3.162}, }
Reference [2]
Michael A. Brilleslyper and Lisbeth E. Schaubroeck, Counting interior roots of trinomials, Math. Mag. 91 (2018), no. 2, 142–150, DOI 10.1080/0025570X.2017.1420332. MR3777918,
Show rawAMSref \bib{bs}{article}{ author={Brilleslyper, Michael A.}, author={Schaubroeck, Lisbeth E.}, title={Counting interior roots of trinomials}, journal={Math. Mag.}, volume={91}, date={2018}, number={2}, pages={142--150}, issn={0025-570X}, review={\MR {3777918}}, doi={10.1080/0025570X.2017.1420332}, }
Reference [3]
Daoud Bshouty, Walter Hengartner, and Tiferet Suez, The exact bound on the number of zeros of harmonic polynomials, J. Anal. Math. 67 (1995), 207–218, DOI 10.1007/BF02787790. MR1383494,
Show rawAMSref \bib{bhs}{article}{ author={Bshouty, Daoud}, author={Hengartner, Walter}, author={Suez, Tiferet}, title={The exact bound on the number of zeros of harmonic polynomials}, journal={J. Anal. Math.}, volume={67}, date={1995}, pages={207--218}, issn={0021-7670}, review={\MR {1383494}}, doi={10.1007/BF02787790}, }
Reference [4]
J. Clunie and T. Sheil-Small, Harmonic univalent functions, Ann. Acad. Sci. Fenn. Ser. A I Math. 9 (1984), 3–25, DOI 10.5186/aasfm.1984.0905. MR752388,
Show rawAMSref \bib{css}{article}{ author={Clunie, J.}, author={Sheil-Small, T.}, title={Harmonic univalent functions}, journal={Ann. Acad. Sci. Fenn. Ser. A I Math.}, volume={9}, date={1984}, pages={3--25}, issn={0066-1953}, review={\MR {752388}}, doi={10.5186/aasfm.1984.0905}, }
Reference [5]
Michael J. Dorff and James S. Rolf, Anamorphosis, mapping problems, and harmonic univalent functions, Explorations in complex analysis, Classr. Res. Mater. Ser., Math. Assoc. America, Washington, DC, 2012, pp. 197–269. MR2963968,
Show rawAMSref \bib{dr}{article}{ author={Dorff, Michael J.}, author={Rolf, James S.}, title={Anamorphosis, mapping problems, and harmonic univalent functions}, conference={ title={Explorations in complex analysis}, }, book={ series={Classr. Res. Mater. Ser.}, publisher={Math. Assoc. America, Washington, DC}, }, date={2012}, pages={197--269}, review={\MR {2963968}}, }
Reference [6]
Peter Duren, Harmonic mappings in the plane, Cambridge Tracts in Mathematics, vol. 156, Cambridge University Press, Cambridge, 2004, DOI 10.1017/CBO9780511546600. MR2048384,
Show rawAMSref \bib{durenbook}{book}{ author={Duren, Peter}, title={Harmonic mappings in the plane}, series={Cambridge Tracts in Mathematics}, volume={156}, publisher={Cambridge University Press, Cambridge}, date={2004}, pages={xii+212}, isbn={0-521-64121-7}, review={\MR {2048384}}, doi={10.1017/CBO9780511546600}, }
Reference [7]
Peter Duren, Walter Hengartner, and Richard S. Laugesen, The argument principle for harmonic functions, Amer. Math. Monthly 103 (1996), no. 5, 411–415, DOI 10.2307/2974933. MR1400723,
Show rawAMSref \bib{dhl}{article}{ author={Duren, Peter}, author={Hengartner, Walter}, author={Laugesen, Richard S.}, title={The argument principle for harmonic functions}, journal={Amer. Math. Monthly}, volume={103}, date={1996}, number={5}, pages={411--415}, issn={0002-9890}, review={\MR {1400723}}, doi={10.2307/2974933}, }
Reference [8]
Russell Howell and David Kyle, Locating trinomial zeros, Involve 11 (2018), no. 4, 711–720, DOI 10.2140/involve.2018.11.711. MR3778921,
Show rawAMSref \bib{hk}{article}{ author={Howell, Russell}, author={Kyle, David}, title={Locating trinomial zeros}, journal={Involve}, volume={11}, date={2018}, number={4}, pages={711--720}, issn={1944-4176}, review={\MR {3778921}}, doi={10.2140/involve.2018.11.711}, }
Reference [9]
Dmitry Khavinson and Grzegorz Świa̧tek, On the number of zeros of certain harmonic polynomials, Proc. Amer. Math. Soc. 131 (2003), no. 2, 409–414, DOI 10.1090/S0002-9939-02-06476-6. MR1933331,
Show rawAMSref \bib{ds}{article}{ author={Khavinson, Dmitry}, author={\'{S}wi\c {a}tek, Grzegorz}, title={On the number of zeros of certain harmonic polynomials}, journal={Proc. Amer. Math. Soc.}, volume={131}, date={2003}, number={2}, pages={409--414}, issn={0002-9939}, review={\MR {1933331}}, doi={10.1090/S0002-9939-02-06476-6}, }
Reference [10]
Marie M. Johnson, Book Review: Analytic Geometry by Charles H. Lehmann, Natl. Math. Mag. 17 (1943), no. 6, 280. MR1570119,
Show rawAMSref \bib{lehmann}{article}{ author={Johnson, Marie M.}, title={Book Review: {\it Analytic Geometry} by Charles H. Lehmann}, journal={Natl. Math. Mag.}, volume={17}, date={1943}, number={6}, pages={280}, issn={1539-5588}, review={\MR {1570119}}, }
Reference [11]
O. Lehto and K. I. Virtanen, Quasiconformal mappings in the plane, 2nd ed., Springer-Verlag, New York-Heidelberg, 1973. Translated from the German by K. W. Lucas; Die Grundlehren der mathematischen Wissenschaften, Band 126. MR0344463,
Show rawAMSref \bib{lehto}{book}{ author={Lehto, O.}, author={Virtanen, K. I.}, title={Quasiconformal mappings in the plane}, edition={2}, note={Translated from the German by K. W. Lucas; Die Grundlehren der mathematischen Wissenschaften, Band 126}, publisher={Springer-Verlag, New York-Heidelberg}, date={1973}, pages={viii+258}, review={\MR {0344463}}, }
Reference [12]
Seung-Yeop Lee, Antonio Lerario, and Erik Lundberg, Remarks on Wilmshurst’s theorem, Indiana Univ. Math. J. 64 (2015), no. 4, 1153–1167, DOI 10.1512/iumj.2015.64.5526. MR3385788,
Show rawAMSref \bib{lll}{article}{ author={Lee, Seung-Yeop}, author={Lerario, Antonio}, author={Lundberg, Erik}, title={Remarks on Wilmshurst's theorem}, journal={Indiana Univ. Math. J.}, volume={64}, date={2015}, number={4}, pages={1153--1167}, issn={0022-2518}, review={\MR {3385788}}, doi={10.1512/iumj.2015.64.5526}, }
Reference [13]
Hans Lewy, On the non-vanishing of the Jacobian in certain one-to-one mappings, Bull. Amer. Math. Soc. 42 (1936), no. 10, 689–692, DOI 10.1090/S0002-9904-1936-06397-4. MR1563404,
Show rawAMSref \bib{lewy}{article}{ author={Lewy, Hans}, title={On the non-vanishing of the Jacobian in certain one-to-one mappings}, journal={Bull. Amer. Math. Soc.}, volume={42}, date={1936}, number={10}, pages={689--692}, issn={0002-9904}, review={\MR {1563404}}, doi={10.1090/S0002-9904-1936-06397-4}, }
Reference [14]
Aaron Melman, Geometry of trinomials, Pacific J. Math. 259 (2012), no. 1, 141–159, DOI 10.2140/pjm.2012.259.141. MR2988487,
Show rawAMSref \bib{mel}{article}{ author={Melman, Aaron}, title={Geometry of trinomials}, journal={Pacific J. Math.}, volume={259}, date={2012}, number={1}, pages={141--159}, issn={0030-8730}, review={\MR {2988487}}, doi={10.2140/pjm.2012.259.141}, }
Reference [15]
Tristan Needham, Visual complex analysis, The Clarendon Press, Oxford University Press, New York, 1997. MR1446490,
Show rawAMSref \bib{needham}{book}{ author={Needham, Tristan}, title={Visual complex analysis}, publisher={The Clarendon Press, Oxford University Press, New York}, date={1997}, pages={xxiv+592}, isbn={0-19-853447-7}, review={\MR {1446490}}, }
Reference [16]
R. Peretz and J. Schmid, Proceedings of the Ashkelon Workshop on Complex Function Theory, (1996), 203–208, Bar-Ilan Univ., Ramat Gan, 1997.
Reference [17]
T. Sheil-Small in Tagesbericht, Mathematisches Forsch. Inst. Oberwolfach, Funktionentheorie, 16-22.2.1992, 19.
Reference [18]
A. S. Wilmshurst, The valence of harmonic polynomials, Proc. Amer. Math. Soc. 126 (1998), no. 7, 2077–2081, DOI 10.1090/S0002-9939-98-04315-9. MR1443416,
Show rawAMSref \bib{wil}{article}{ author={Wilmshurst, A. S.}, title={The valence of harmonic polynomials}, journal={Proc. Amer. Math. Soc.}, volume={126}, date={1998}, number={7}, pages={2077--2081}, issn={0002-9939}, review={\MR {1443416}}, doi={10.1090/S0002-9939-98-04315-9}, }

Article Information

MSC 2010
Primary: 30C15 (Zeros of polynomials, rational functions, and other analytic functions)
Author Information
Michael Brilleslyper
Department of Mathematical Sciences, United States Air Force Academy, USAF Academy, Colorado 80840
mike.brilleslyper@usafa.edu
MathSciNet
Jennifer Brooks
Department of Mathematics, Brigham Young University, Provo, Utah 84602
jbrooks@mathematics.byu.edu
MathSciNet
Michael Dorff
Department of Mathematics, Brigham Young University, Provo, Utah 84602
mdorff@math.byu.edu
ORCID
MathSciNet
Russell Howell
Department of Mathematics and Computer Science, Westmont College, Santa Barbara, California 93108
howell@westmont.edu
ORCID
MathSciNet
Lisbeth Schaubroeck
Department of Mathematical Sciences, United States Air Force Academy, USAF Academy, Colorado 80840
beth.schaubroeck@usafa.edu
MathSciNet
Communicated by
Jeremy Tyson
Journal Information
Proceedings of the American Mathematical Society, Series B, Volume 7, Issue 7, ISSN 2330-1511, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2020 by the authors under Creative Commons Attribution 3.0 License (CC BY 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/bproc/51
  • MathSciNet Review: 4127911
  • Show rawAMSref \bib{4127911}{article}{ author={Brilleslyper, Michael}, author={Brooks, Jennifer}, author={Dorff, Michael}, author={Howell, Russell}, author={Schaubroeck, Lisbeth}, title={Zeros of a one-parameter family of harmonic trinomials}, journal={Proc. Amer. Math. Soc. Ser. B}, volume={7}, number={7}, date={2020}, pages={82-90}, issn={2330-1511}, review={4127911}, doi={10.1090/bproc/51}, }

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.