1. Trang chủ
  2. » Khoa Học Tự Nhiên

THE AMERICAN MATHEMATICAL MONTHLY

701 428 1

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

THÔNG TIN TÀI LIỆU

Thông tin cơ bản

Tiêu đề The American Mathematical Monthly Volume 117, Number 1 (January 2010)
Tác giả Aaron Abrams, Skip Garibaldi, Deborah E. Berg, Serguei Norine, Francis Edward Su, Robin Thomas, Paul Wollan, Christopher S. Hardin, Barry Lewis
Trường học Amherst College
Chuyên ngành Mathematics
Thể loại journal issue
Năm xuất bản 2010
Thành phố Amherst
Định dạng
Số trang 701
Dung lượng 14,95 MB

Các công cụ chuyển đổi và chỉnh sửa cho tài liệu này

Nội dung

Tạp chí toán học AMM của Mỹ

Trang 1

THE AMERICAN MATHEMATICAL

Serguei Norine

Francis Edward Su

Robin Thomas

Paul Wollan

NOTES

in Mathematics and Science Edited by

Chandler Davis, Marjorie Wikler Senechal, and Jan Zwicky

AN OFFICIAL PUBLICATION OF THE MATHEMATICAL ASSOCIATION OF AMERICA

Trang 3

THE AMERICAN MATHEMATICAL

EDITOR Daniel J Velleman Amherst College ASSOCIATE EDITORS William Adkins

Louisiana State University

National Science Foundation Joel W Robbin

University of Wisconsin, Madison Rachel Roberts

Washington University, St Louis Judith Roitman

University of Kansas, Lawrence Edward Scheinerman

Johns Hopkins University Abe Shenitzer

York University Karen E Smith University of Michigan, Ann Arbor Susan G Staples

Texas Christian University John Stillwell

University of San Francisco Dennis Stowe

Idaho State University, Pocatello Francis Edward Su

Harvey Mudd College Serge Tabachnikov Pennsylvania State University Daniel Ullman

George Washington University Gerard Venema

Calvin College Douglas B West University of Illinois, Urbana-Champaign

EDITORIAL ASSISTANT Nancy R Board

Trang 4

NOTICE TO AUTHORS

The MONTHLY publishes articles, as well as notes

and other features, about mathematics and the

pro-fession Its readers span a broad spectrum of

math-ematical interests, and include professional

mathe-maticians as well as students of mathematics at all

collegiate levels Authors are invited to submit articles

and notes that bring interesting mathematical ideas

to a wide audience of MONTHLY readers.

The MONTHLY’s readers expect a high standard of

exposition; they expect articles to inform, stimulate,

challenge, enlighten, and even entertain MONTHLY

articles are meant to be read, enjoyed, and

dis-cussed, rather than just archived Articles may be

expositions of old or new results, historical or

bi-ographical essays, speculations or definitive

treat-ments, broad developtreat-ments, or explorations of a

sin-gle application Novelty and generality are far less

im-portant than clarity of exposition and broad appeal.

Appropriate figures, diagrams, and photographs are

encouraged.

Notes are short, sharply focused, and possibly

infor-mal They are often gems that provide a new proof

of an old theorem, a novel presentation of a familiar

theme, or a lively discussion of a single issue.

Articles and notes should be sent to the Editor:

For an initial submission, please send a pdf file as

an email attachment to: mathmonthly@amherst.edu.

(Pdf is the only electronic file format we accept.)

Please put “Submission to the Monthly” in the

sub-ject line, and include the title of the paper and

the name and postal address of the corresponding

author in the body of your email If submitting more

than one paper, send each in a separate email In

lieu of a pdf, an author may submit a single paper

copy of the manuscript, printed on only one side of

the paper Manuscript pages should be numbered,

and left and right margins should be at least one

inch wide Authors who use L A TEX are urged to use

article.sty, or a similar generic style, and its

stan-dard environments with no custom formatting See

recent articles in the MONTHLY for the style of

ci-tations for journal articles and books Follow the

link to Electronic Publication Information for authors

at http://www.maa.org/pubs/monthly.html for

informa-tion about figures and files as well as general editorial

guidelines.

Letters to the Editor on any topic are invited

Com-ments, criticisms, and suggestions for making the

MONTHLY more lively, entertaining, and informative

are welcome.

The online MONTHLY archive at www.jstor.org is a

valuable resource for both authors and readers; it

may be searched online in a variety of ways for any

specified keyword(s) MAA members whose

institu-tions do not provide JSTOR access may obtain

indi-vidual access for a modest annual fee; call

800-331-1622.

See the MONTHLY section of MAA Online for current

information such as contents of issues and

descrip-tive summaries of forthcoming articles:

http://www.maa.org/

Proposed problems or solutions should be sent to: DOUG HENSLEY, MONTHLY Problems Department of Mathematics

Texas A&M University

3368 TAMU College Station, TX 77843-3368

In lieu of duplicate hardcopy, authors may submit pdfs to monthlyproblems@math.tamu.edu.

Advertising Correspondence:

MAA Advertising

1529 Eighteenth St NW Washington DC 20036 Phone: (866) 821-1221 Fax: (866) 387-1208 E-mail: advertising@maa.org Further advertising information can be found online

at www.maa.org Change of address, missing issue inquiries, and other subscription correspondence:

MAA Service Center, maahq@maa.org All at the address:

The Mathematical Association of America

1529 Eighteenth Street, N.W.

Washington, DC 20036 Recent copies of the MONTHLY are available for pur- chase through the MAA Service Center.

maahq@maa.org, 1-800-331-1622 Microfilm Editions: University Microfilms Interna- tional, Serial Bid coordinator, 300 North Zeeb Road, Ann Arbor, MI 48106.

The AMERICAN MATHEMATICAL MONTHLY (ISSN 0002-9890) is published monthly except bimonthly June-July and August-September by the Mathemati- cal Association of America at 1529 Eighteenth Street, N.W., Washington, DC 20036 and Montpelier, VT, and copyrighted by the Mathematical Association

of America (Incorporated), 2010, including rights to this journal issue as a whole and, except where otherwise noted, rights to each individual contribu- tion Permission to make copies of individual ar- ticles, in paper or electronic form, including post- ing on personal and class web pages, for educa- tional and scientific use is granted without fee pro- vided that copies are not made or distributed for profit or commercial advantage and that copies bear the following copyright notice: [Copyright the Math- ematical Association of America 2010 All rights re- served.] Abstracting, with credit, is permitted To copy otherwise, or to republish, requires specific permis- sion of the MAA’s Director of Publications and pos- sibly a fee Periodicals postage paid at Washing- ton, DC, and additional mailing offices.Postmaster: Send address changes to the American Mathemat- ical Monthly, Membership/Subscription Department, MAA, 1529 Eighteenth Street, N.W., Washington,

DC, 20036-1385.

Trang 5

Finding Good Bets in the Lottery,

and Why You Shouldn’t Take Them

Aaron Abrams and Skip Garibaldi

Everybody knows that the lottery is a bad investment But do you know why? How doyou know?

For most lotteries, the obvious answer is obviously correct: lottery operators arerunning a business, and we can assume they have set up the game so that they makemoney If they make money, they must be paying out less than they are taking in;

so on average, the ticket buyer loses money This reasoning applies, for example, to

the policy games formerly run by organized crime described in [15] and [18], and

to the (essentially identical) Cash 3 and Cash 4 games currently offered in the state

of Georgia, where the authors reside This reasoning also applies to Las Vegas–stylegambling (How do you think the Luxor can afford to keep its spotlight lit?)

However, the question becomes less trivial for games with rolling jackpots, like

Mega Millions (currently played in 12 of the 50 U.S states), Powerball (played in 30states), and various U.S state lotteries In these games, if no player wins the largest

prize (the jackpot) in any particular drawing, then that money is “rolled over” and

increased for the next drawing On average the operators of the game still make money,but for a particular drawing, one can imagine that a sufficiently large jackpot wouldgive a lottery ticket a positive average return (even though the probability of winningthe jackpot with a single ticket remains extremely small) Indeed, for any particulardrawing, it is easy enough to calculate the expected rate of return using the formula

in (4.5) This has been done in the literature for lots of drawings (see, e.g., [20]), and,

sure enough, sometimes the expected rate of return is positive In this situation, why isthe lottery a bad investment?

Seeking an answer to this question, we began by studying historical lottery data Indoing so, we were surprised both by which lotteries offered the good bets, and also

by just how good they can be We almost thought we should invest in the lottery!

So we were faced with several questions; for example, are there any rules of thumb

to help pick out the drawings with good rates of return? One jackpot winner saidshe only bought lottery tickets when the announced jackpot was at least $100 million

[26] Is this a good idea? (Or perhaps a modified version, replacing the threshold with

something less arbitrary?) Sometimes the announced jackpots of these games are trulyenormous, such as on March 9, 2007, when Mega Millions announced a $390 millionprize Would it have been a good idea to buy a ticket for that drawing? And our realquestion, in general, is the following: on the occasions that the lottery offers a positiverate of return, is a lottery ticket ever a good investment? And how can we tell? In thispaper we document our findings Using elementary mathematics and economics, wecan give satisfying answers to these questions

We should come clean here and admit that to this point we have been deliberatelyconflating several notions By a “good bet” (for instance in the title of this paper) wemean any wager with a positive rate of return This is a mathematical quantity which

is easily computed A “good investment” is harder to define, and must take into count risk This is where things really get interesting, because as any undergraduateeconomics major knows, mathematics alone does not provide the tools to determine

ac-doi:10.4169/000298910X474952

Trang 6

when a good bet is a good investment (although a bad bet is always a bad investment!).

To address this issue we therefore leave the domain of mathematics and enter a cussion of some basic economic theory, which, in Part III of the paper, succeeds inanswering our questions (hence the second part of the title) And by the way, a “goodidea” is even less formal: independently of your financial goals and strategies, youmight enjoy playing the lottery for a variety of reasons We’re not going to try to stopyou

dis-To get started, we build a mathematical model of a lottery drawing Part I of this per (§§ 1–3) describes the model in detail: it has three parameters ( f , F, t) that depend only on the lottery and not on a particular drawing, and two parameters (N , J) that vary from drawing to drawing Here N is the total ticket sales and J is the size of the jackpot (The reader interested in a particular lottery can easily determine f , F, and

pa-t ) The benefipa-t of pa-the general model, of course, is pa-thapa-t ipa-t allows us pa-to prove pa-theorems

about general lotteries The parameters are free enough that the results apply to MegaMillions, Powerball, and many other smaller lotteries

In Part II (§§4–8) we use elementary calculus to derive criteria for determining,without too much effort, whether a given drawing is a good bet We show, roughlyspeaking, that drawings with “small” ticket sales (relative to the jackpot; the measure-

ment we use is N /J, which should be less than 1/5) offer positive rates of return, once

the jackpot exceeds a certain easily-computed threshold Lotto Texas is an example ofsuch a lottery On the other hand, drawings with “large” ticket sales (again, this means

N /J is larger than a certain cutoff, which is slightly larger than 1) will always have

negative rates of return As it happens, Mega Millions and Powerball fall into this gory; in particular, no drawing of either of these two lotteries has ever been a good bet,including the aforementioned $390 million jackpot Moreover, based on these consid-erations we argue in Section 8 that Mega Millions and Powerball drawings are likely

cate-to always be bad bets in the future also

With this information in hand, we focus on those drawings that have positive pected rates of return, i.e., the good bets, and we ask, from an economic point of view,whether they can ever present a good investment If you buy a ticket, of course, youwill most likely lose your dollar; on the other hand, there is a small chance that youwill win big Indeed, this is the nature of investing (and gambling): every interestinginvestment offers the potential of gain alongside the risk of loss If you view the lot-tery as a game, like playing roulette, then you are probably playing for fun and youare both willing and expecting to lose your dollar But what if you really want to makemoney? Can you do it with the lottery? More generally, how do you compare invest-ments whose expected rates of return and risks differ?

ex-In Part III of the paper (§§9–11) we discuss basic portfolio theory, a branch ofeconomics that gives a concrete, quantitative answer to exactly this question Portfoliotheory is part of a standard undergraduate economics curriculum, but it is not so wellknown to mathematicians Applying portfolio theory to the lottery, we find, as onemight expect, that even when the returns are favorable, the risk of a lottery ticket is

so large that an optimal investment portfolio will allocate a negligible fraction of its

assets to lottery tickets Our conclusion, then, is unsurprising; to quote the movie War

Games, “the only winning move is not to play.”a

You might respond: “So what? I already knew that buying a lottery ticket was a badinvestment.” And maybe you did But we thought we knew it too, until we discoveredthe fantastic expected rates of return offered by certain lottery drawings! The point

we want to make here is that if you want to actually prove that the lottery is a bad

a How about a nice game of chess?

Trang 7

investment, the existence of good bets shows that mathematics is not enough It takeseconomics in cooperation with mathematics to ultimately validate our intuition.

Further Reading The lotteries described here are all modern variations on a lottery invented in Genoa in the 1600s, allegedly to select new senators [1] The Genoese-style

lottery became very popular in Europe, leading to interest by mathematicians including

Euler; see, e.g., [9] or [2] The papers [21] and [35] survey modern U.S lotteries from

an economist’s perspective The book [34] gives a treatment for a general audience.

The conclusion of Part III of the present paper—that even with a very good expectedrate of return, lotteries are still too risky to make good investments—has of course

been observed before by economists; see [17] Whereas we compare the lottery to other investments via portfolio theory, the paper [17] analyzes a lottery ticket as an investment in isolation, using the Kelly criterion (described, e.g., in [28] or [32]) to

decide whether or not to invest They also conclude that one shouldn’t invest in thelottery, but for a different reason than ours: they argue that investing in the lotteryusing a strategy based on the Kelly criterion, even under favorable conditions, is likely

to take millions of years to produce positive returns The mathematics required fortheir analysis is more sophisticated than the undergraduate-level material used in ours

P ART I THE SETUP: MODELING A LOTTERY

1 Mega Millions and Powerball The Mega Millions and Powerball lotteries are

similar in that in both, a player purchasing a $1 ticket selects 5 distinct “main” numbers(from 1 to 56 in Mega Millions and 1 to 55 in Powerball) and 1 “extra” number (from

1 to 46 in Mega Millions and 1 to 42 in Powerball) This extra number is not related tothe main numbers, so, e.g., the sequence

main= 4, 8, 15, 16, 23 and extra = 15

denotes a valid ticket in either lottery The number of possible distinct tickets is56

5

46for Mega Millions and55

fixed, because their values are fixed (In this paper, we treat a slightly simplified version

of the Powerball game offered from August 28, 2005 through the end of 2008 Theactual game allowed the player the option of buying a $2 ticket that had larger fixedprizes Also, in the event of a record-breaking jackpot, some of the fixed prizes werealso increased by a variable amount We ignore both of these possibilities The rulesfor Mega Millions also vary slightly from state to state,band we take the simplest andmost popular version here.)

The payouts listed in our table for the two largest fixed prizes are slightly differentfrom those listed on the lottery websites, in that we have deducted federal taxes Cur-rently, gambling winnings over $600 are subject to federal income tax, and winnings

over $5000 are withheld at a rate of 25%; see [14] or [4] Since income tax rates vary

from gambler to gambler, we use 25% as an estimate of the tax rate.cFor example, the

b Most notably, in California all prizes are pari-mutuel.

c We guess that most people who win the lottery will pay at least 25% in taxes For anyone who pays more,

the estimates we give of the jackpot value J for any particular drawing should be decreased accordingly This

kind of change strengthens our final conclusion—namely, that buying lottery tickets is a poor investment.

Trang 8

Table 1 Prizes for Mega Millions and Powerball A ticket costs $1 Payouts for the

5/5 and 4/5 + extra prizes have been reduced by 25% to approximate taxes.

5/5 + extra jackpot 1 jackpot 1

Jersey [24, p 19] and New Hampshire respectively), we ignore state taxes for these

lotteries

2 Lotteries with Other Pari-mutuel Prizes In addition to Mega Millions or

Power-ball, some states offer their own lotteries with rolling jackpots Here we describe theTexas (“Lotto Texas”) and New Jersey (“Pick 6”) games In both, a ticket costs $1 andconsists of 6 numbers (1–49 for New Jersey and 1–54 for Texas) For matching 3 ofthe 6 winning numbers, the player wins a fixed prize of $3

All tickets that match 4 of the 6 winning numbers split a pot of.05N (NJ) or 033N (TX) dollars, where N is the total amount of sales for that drawing (As tickets cost $1,

as a number, N is the same as the total number of tickets sold.) The prize for matching

5 of the 6 winning numbers is similar; such tickets split a pot of.055N (NJ) or 0223N

(TX); these prizes are typically around $2000, so we deduct 25% in taxes from them

as in the previous section, resulting in.0413N for New Jersey and 0167N for Texas.

(Deducting this 25% makes no difference to any of our conclusions, it only slightlychanges a few numbers along the way.) Finally, the tickets that match all 6 of the 6winning numbers split the jackpot

How did we find these rates? For New Jersey, they are on the state lottery website.Otherwise, you can approximate them from knowing—for a single past drawing—theprize won by each ticket that matched 4 or 5 of the winning numbers, the number of

tickets sold that matched 4 or 5 of the winning numbers, and the total sales N for that

drawing (In the case of Texas, these numbers can be found on the state lottery website,

except for total sales, which is on the website of a third party [23].) The resulting

estimates may not be precisely correct, because the lottery operators typically roundthe prize given to each ticket holder to the nearest dollar

As a matter of convenience, we refer to the prize for matching 3 of 6 as fixed, the prizes for matching 4 or 5 of 6 as pari-mutuel, and the prize for matching 6 of 6 as the jackpot (Strictly speaking, this is an abuse of language, because the jackpot is also

pari-mutuel in the usual sense of the word.)

Trang 9

3 The General Model We want a mathematical model that includes Mega Millions,

Powerball, and the New Jersey and Texas lotteries described in the preceding section

We model an individual drawing of such a lottery, and write N for the total ticket sales,

an amount of money Let t be the number of distinct possible tickets, of which:

1 of the possible distinct tickets wins a share of the (positive) jackpot J More

pre-cisely, ifw copies of this 1 ticket are sold, then each ticket-holder receives J/w Note that the a i , the r i , and the various t’s depend only on the setup of the lottery, whereas N and J vary from drawing to drawing Also, we mention a few technical points The prizes a i , the number N , and the jackpot J are denominated in units of

“price of a ticket.” For all four of our example lotteries, the tickets cost $1, so, for

example, the amounts listed in Table 1 are the a i’s—one just drops the dollar sign

Furthermore, the prizes are the actual amount the player receives We assume that

taxes have already been deducted from these amounts, at whatever rate such winningswould be taxed (In this way, we avoid having to include tax in our formulas.) Jackpotwinners typically have the option of receiving their winnings as a lump sum or as an

annuity; see, e.g., [31] for an explanation of the differences We take J to be the

after-tax value of the lump sum, or—what is the essentially the same—the present value

(after tax) of the annuity Note that this J is far smaller than the jackpot amounts

an-nounced by lottery operators, which are usually totals of the pre-tax annuity payments.Some comparisons are shown in Table 3A

Table 3 A Comparison of annuity and lump sum jackpot amounts for some lottery drawings.

The value of J is the lump sum minus tax, which we assume to be 25% The letter ‘m’ denotes

millions of dollars.

Annuity jackpot (pre-tax)

Lump sum jackpot (pre-tax) J (estimated)

We assume that the player knows J After all, the pre-tax value of the annuitized

jackpot is announced publicly in advance of the drawing, and from it one can estimate

J For Mega Millions and Powerball, the lottery websites also list the pre-tax value of

the cash jackpot, so the player only needs to consider taxes

Statistics In order to analyze this model, we focus on a few statistics f , F, and J0

deduced from the data above These numbers depend only on the lottery itself (e.g.,Mega Millions), and not on a particular drawing

We define f to be the cost of a ticket less the expected winnings from fixed prizes,

Trang 10

This number is approximately the proportion of lottery sales that go to the jackpot, thepari-mutuel prizes, and “overhead” (i.e., the cost of lottery operations plus vigorish;around 45% of total sales for the example lotteries in this paper) It is not quite the

same, because we have deducted income taxes from the amounts a i Because the a i are positive, we have f ≤ 1

We call this quantity the jackpot cutoff, for reasons which will become apparent in

Section 5 Table 3B lists these numbers for our four example lotteries We assumed

that some ticket can win the jackpot, so t ≥ 1 and consequently J0 > 0.

Table 3 B Some statistics for our example lotteries that hold for all drawings.

P ART II TO BET OR NOT TO BET: ANALYZING THE RATE OF RETURN

4 Expected Rate of Return Using the model described in Part I, we now calculate

the expected rate of return (eRoR) on a lottery ticket, assuming that a total of N tickets

are sold The eRoR is

(eRoR) = −

cost ofticket

+

expected winningsfrom fixed prizes



where all the terms on the right are measured in units of “cost of one ticket.” The

parameter f defined in (3.1) is the negative of the first two terms.

We focus on one ticket: yours We will assume that the particular numbers on theother tickets are chosen randomly See 4.10 below for more on this hypothesis Withthis assumption, the probability that your ticket is a jackpot winner together withw − 1

of the other tickets is given by the binomial distribution:



N − 1

w − 1

 1

Trang 11

In this case the amount you win is J /w We therefore define

and now the expected amount won from the jackpot is J s (1/t, N) Combining this

with a similar computation for the pari-mutuel prizes and with the preceding graph, we obtain the formula:

t , N



where p i := tpari

i /t is the probability of winning the ith pari-mutuel prize and (d is the

number of pari-mutuel prizes)

The domain of the function s can be extended to allow N to be real (and not just

integral); one can do this the same way one ordinarily treats binomial coefficients in

calculus, or equivalently by using the closed form expression for s given in Proposition

4.4

In the rest of this section we use first-year calculus to prove some basic facts about

the function s (p, x) An editor pointed out to us that our proofs could be shortened

if we approximated the binomial distribution (4.2) by a Poisson distribution (with rameterλ = N/t) Indeed this approximation is extremely good for the relevant values

pa-of N and t; however, since we are able to obtain the results we need from the exact

formula (4.2), we will stick with this expression in order to keep the exposition contained

self-Proposition 4.4 For 0 < p < 1 and N > 0,

Table 4 Expected rate of return for some specific drawings, calculated using (4.5).

Trang 12

Proof of Prop 4.4 The binomial theorem gives an equality of functions of two



x k+1y n −k

The proposition follows by plugging in x = p, y = 1 − p, N = n + 1, and w = k + 1.

We will apply the next lemma repeatedly in what follows

Lemma 4.6 If 0 < c < 1 then 1 −1

c − ln c < 0.

Proof It is a common calculus exercise to show that 1 + z < e z for all z > 0 (e.g., by

using a power series) This implies ln(1 + z) < z or, shifting the variable, ln z < z − 1 for all z > 1 Applying this with z = 1/c completes the proof.

Lemma 4.7 Fix b ∈ (0, 1).

(1) The function

h (x, y) := 1− b x y

x satisfies h x := ∂h/∂x < 0 and h y := ∂h/∂y > 0 for x, y > 0.

(2) For every z > 0, the level set {(x, y) | h(x, y) = z} intersects the first quadrant

in the graph of a smooth, positive, increasing, concave up function defined on the interval (0, 1/z).

Proof We first prove (1) The partial with respect to y is easy: h y = −b x yln(b) > 0 For the partial with respect to x, we have:

(with c = b x y ) Thus h x < 0.

Trang 13

To prove (2), we fix z > 0 and solve the equation h(x, y) = z for y, obtaining

To establish the concavity we differentiate again and rearrange (considerably), taining

is positive, where c := 1 − xz is between 0 and 1.

To show that (4.8) is positive, note first that it is decreasing in c (This is because the derivative is 4c (1 −1

c − ln c), which is negative by Lemma 4.6.) Now, at c = 1, the

value of (4.8) is 0, so for 0< c < 1 we deduce that (4.8) is positive Hence d2y /dx2ispositive, and the level curve is concave up, as claimed

Corollary 4.9 Suppose that 0 < p < 1 For x ≥ 0,

(1) the function x → s(p, x) decreases from − ln(1 − p) to 0.

(2) the function x → x s(p, x) increases from 0 to 1.

Proof.

(1) Lemma 4.7 implies that s (p, x) is decreasing, by setting y = 1 and b = 1 − p Its limit as x → ∞ is obvious because 0 < 1 − p < 1, and its limit as x → 0

can be obtained via l’Hospital’s rule

(2) Proposition 4.4 gives that x s (p, x) is equal to 1 − (1 − p) x, whence it is creasing and has the claimed limits because 0< 1 − p < 1.

in-4.10 Unpopular Numbers Throughout this paper, we are assuming that the other

lottery players select their tickets randomly This is not strictly true: in a typicalU.S lottery drawing, only 70 to 80 percent of the tickets sold are “quick picks,” i.e.,

tickets whose numbers are picked randomly by computer [27, 21], whereas the

oth-ers have numboth-ers chosen by the player Ample evidence indicates that player-chosennumbers are not evenly distributed amongst the possible choices, leading to a strategy:

by playing “unpopular” numbers, you won’t improve your chances of winning anyparticular prize, but if you do win the jackpot, your chance of sharing it decreases

This raises your expected return See, e.g., [6], [13], [34], or the survey in§4 of [12].

Trang 14

The interested reader can adjust the results of this paper to account for this strategy.One way to proceed is to view (4.3) as a lower bound on the eRoR, and then compute

an upper bound by imagining the extreme scenario in which one plays numbers thatare not hand-picked by any other player In this case any sharing of the jackpot willresult from quick picks, which we assume are used by at least 70% of all players

Therefore one can replace N by N = 0.7N in (4.2) and hence in (4.3) The resulting

value of eRoR will be an upper bound on the returns available using the approach ofchoosing unpopular numbers

5 The Jackpot Cutoff J0 With the results of the preceding section in hand, it is

easy to see that the rate of return on “nearly all” lottery drawings is negative How?

We prove an upper bound on the eRoR (4.5)

In (4.5), the terms 1− (1 − p i ) N are at most 1 So we find:

easy-to-ber 2008) where the jackpot J was at least J0 There are only 3 examples out of about

700 drawings, so this event is uncommon We also include one Lotto Texas drawing

with J > J0 (This drawing was preceded by a streak of several such drawings inwhich no one won the jackpot; the eRoR increased until someone won the April 7,

2007 drawing.) During 2007, the New Jersey lottery had no drawings with J > J0.Table 4 also includes, for each drawing, an estimate of the number of tickets soldfor that drawing The precise numbers are not publicized by the lottery operators In

the case of Powerball, we estimated N by using the number of tickets that won a prize,

as reported in press releases For Mega Millions, we used the number of tickets thatwon either of the two smallest fixed prizes, as announced on the lottery website For

Lotto Texas, the number of tickets sold is from [23].

6 Break-even Curves At this point, for any particular drawing, we are able to

ap-ply (4.5) to compute the expected rate of return But this quickly becomes tiresome;

Trang 15

our goal in this section is to give criteria that—in “most” cases—can be used to

deter-mine the sign of the expected rate of return, i.e., whether or not the drawing is a good

bet These criteria will be easy to check, making them applicable to large classes ofdrawings all at once

Recall that for a given lottery, the numbers f , r i , p i , and t (and therefore F and

J0) are constants depending on the setup of the lottery; what changes from drawing to

drawing are the values of J and N Thus J and N are the parameters which control the eRoR for any particular drawing In light of (5.2), it makes sense to normalize J

by dividing it by the jackpot cutoff J0; equation (5.2) says that if J /J0 < 1 then the eRoR is negative It is helpful to normalize N as well; it turns out that a good way

to do this is to divide by J , as the quantity N /J plays a decisive role in our analysis.

Thus we think of ticket sales as being “large” or “small” only in relation to the size ofthe jackpot We therefore use the variables

For any particular lottery, we can plug in the actual values for the parameters f , r i,

p i , p, and J0 and plot the level set{(eRoR) = 0} in the first quadrant of the (x, plane We call this level set the break-even curve, because drawings lying on this curve

y)-have expected rate of return zero This curve is interesting for the gambler because itseparates the region of the(x, y)-plane consisting of drawings with positive rates of

return from the region of those with negative rates of return

Proposition 6.2 For every lottery as in § 3, the break-even curve is the graph of a

smooth, positive function (x) with domain (0, 1/F) If the lottery has no pari-mutuel

prizes, then (x) is increasing and concave up.

Proof For a fixed positive x, we can use (6.1) to view the eRoR as a (smooth) function

of one variable y At y = 0, we have (eRoR) = − f < 0 Looking at (6.1), we see that the nonconstant terms all have positive partial derivatives with respect to y, so eRoR

is a continuous, strictly increasing function of y Further,

y > 0 such that (eRoR) = 0 Moreover for x ≥ 1/F the eRoR is always negative This

proves that the break-even curve is the graph of a positive function(x) with domain (0, 1/F).

Now, observe that for fixed y > 0, it is in general unclear whether eRoR is ing in x One expects it to decrease, because x is proportional to ticket sales However,

increas-in (6.1), the terms increas-in the sum (which are typically small) are increas-increasincreas-ing increas-in x whereas only the last term is decreasing in x (by Lemma 4.7) If there are no pari-mutuel prizes,

Trang 16

then the sum disappears, and the eRoR is decreasing in x Specifically, in this case the

graph of(x) is the level curve of the function h(x, y) = (1 − b x y )/x at the value f , where b = (1 − p) J0 By Lemma 4.7,(x) is smooth, increasing, and concave up.

Nevertheless, even when there are pari-mutuel prizes, the function (x) is still smooth This is because, as we showed in the first paragraph of this proof, the y-

derivative of eRoR is always positive (for 0< x < 1/F), so we can invoke the implicit

function theorem Since eRoR is smooth as a function of two variables, its level sets

are graphs of smooth functions of x on this interval The break-even curve is one such

level set

The break-even curves for Mega Millions, Powerball, Lotto Texas, and New JerseyPick 6 are the lighter curves in Figure 6A (Note that they are so similar that it is hard

to tell that there are four of them!)

The two bold curves in Figure 6Aare the level curves

Their significance is spelled out in the following theorem, which is our core result We

consider a lottery to be major if there are at least 500 distinct tickets; one can see from

the proofs that this particular bound is almost arbitrary

0.5 1.0 1.5

2.0

J / J0

N / J

Figure 6 A Illustration of Theorem 6.5 The break-even curve for any lottery must lie in the region between the

two bold curves The lighter curves between the two bold curves are the break-even curves for Mega Millions, Powerball, Lotto Texas, and New Jersey Pick 6 Drawings in the region marked − have a negative eRoR and those in the region marked + have a positive eRoR.

Theorem 6.5 For any major lottery with F ≥ 0.8 (i.e., that roughly speaking pays

out on average less than 20% of its revenue in prizes other than the jackpot) we have:

Trang 17

(a) The break-even curve lies in the region between the curves U and L and to the

right of the y-axis.

(b) Any drawing with (x, y) above the break-even curve has positive eRoR.

(c) Any drawing with (x, y) below the break-even curve has negative eRoR.

In particular, for any drawing of any such lottery, if (x, y) is above U then the eRoR

is positive, and if (x, y) is below or to the right of L then the eRoR is negative.

The hypotheses of the theorem are fairly weak Being major means that the lotteryhas at least 500 possible distinct tickets; our example lotteries all have well over a

million distinct tickets All of our example lotteries have F ≥ 0.82.

We need the following lemma from calculus

Lemma 6.6 The function g (t) = (1 − 1/t) t is increasing for t > 1, and the limit is

1/e.

Proof The limit is well known We prove that g is increasing, which is an exercise in

first-year calculus.dIn order to take the derivative we first take the logarithm:

We want to show that g (t) > 0 Since the denominator on the left, g(t), is positive, it

suffices to show that the right side is positive To do this we rearrange the logarithm as

interval of integration

We can now prove the theorem

Proof of Theorem 6.5 Parts (b) and (c) follow from the proof of Proposition 6.2: for a

fixed x, the eRoR is an increasing function of y.

To prove (a), we first claim that for any c > 0,

1− 0.45 c < cJ0s

1

Trang 18

Recall that J0= Ft We write out s using Proposition 4.4:

c J0s

1

For the lower bound in (6.8), Lemma 6.6 implies that(1 − 1/t) tis at most 1/e Putting

these together, we find:

1− e −cF < cJ0s

1

Looking at equation (4.5) for the eRoR, we observe that the pari-mutuel rates r iare

nonnegative, so for a lower bound we can take p i = 0 for all i Combined with the

upper bound (5.1), we have:

− f + J s

1

t , N



≤ (eRoR) ≤ −F + J s

1

t , N



Replacing N with x y J0and J with x y J0/x and applying (6.8) with c = xy gives:

−1 +1− 0.45 x y

x < (eRoR) < −0.8 +1− 0.36 x y

By Lemma 4.7, the partial derivatives of the upper and lower bounds in (6.9) are

negative with respect to x and positive with respect to y This implies (a) as well as

the last sentence of the theorem

We can enlarge the negative region in Figure 6Asomewhat by incorporating (5.2),

which says that the eRoR is negative for any drawing with y = J/J0 < 1 In fact (5.2) implies that the break-even curve for a lottery will not intersect the line y = 1 except

possibly when no other tickets are sold, i.e., when N = 1 or equivalently x = 1/J ≈ 0.

The result is Figure 6B

Figure 6B includes several data points for actual drawings which have occurred.The four solid dots represent the four drawings from Table 4 The circles in the bot-tom half of the figure are a few typical Powerball and Mega Millions drawings Inall the drawings that we examined, the only ones we found in the inconclusive region(between the bold curves) were some of those leading up to the positive Lotto Texasdrawing plotted in the figure

7 Examples We now give two concrete illustrations of Theorem 6.5 Let’s start with

the good news

7.1 Small Ticket Sales The point (0.2, 1.4) (approximately) is on the curve U

de-fined in (6.3) Any drawing of any lottery satisfying

Trang 19

+ + +

2.0

J / J0

N / J

Figure 6 B Refinement of Figure 6A Drawings in the regions marked with +’s have positive eRoR and those

in regions marked with −’s have negative eRoR.

(1) N < 0.2 J and

(2) J > 1.4 J0

is above U and so will have positive eRoR by Theorem 6.5 (The points satisfying (1)

and (2) make up the small shaded rectangle on the left side of Figure 7.) We chose

to look at this point because some state lotteries, such as Lotto Texas, tend to satisfy

(1) every week So to find a positive eRoR, this is the place to look: just wait until J

reaches the threshold (2)

+ + +

Trang 20

7.2 Large Ticket Sales On the other hand, the point(1.12, 2) is on the curve L

de-fined in (6.4); therefore any drawing of any lottery with

(1) N > 1.12 J and

(2) J < 2 J0

lies below or to the right of L and will have a negative eRoR These drawings make

up the large rectangular region on the right side of Figure 7 Here we have chosen tofocus on this particular rectangle for two reasons Mega Millions and Powerball tend

to have large ticket sales (relative to J ); specifically, N /J has exceeded 1.12 every time J has exceeded J0 Moreover, no drawing of any lottery we are aware of has evercome close to violating (2) In fact, for lotteries with large ticket sales (such as Mega

Millions and Powerball), the largest value of J /J0 we have observed is about 1.19,

in the case of the Mega Millions drawing in Table 4 Thus no past drawing of MegaMillions or Powerball has ever offered a positive eRoR

Of course, if we are interested in Mega Millions and Powerball specifically, then

we may obtain stronger results by using their actual break-even curves, rather than

the bound L We will do this in the next section to argue that in all likelihood these

lotteries will never offer a good bet

In each of these examples, our choice of region (encoded in the hypotheses (1) and(2)) is somewhat arbitrary The reader who prefers a different rectangular region can

easily cook one up: just choose a point on U or L to be the (lower right or upper left)

corner of the rectangle and apply Theorem 6.5 At the cost of slightly more cated (but still linear) hypotheses, one could prove something about various triangularregions as well

compli-8 Mega Millions and Powerball Mega Millions and Powerball fall under the Large

Ticket Sales example (7.2) of the previous section, and indeed they have never offered

a positive eRoR We now argue that in all likelihood, no future drawing of either of

these lotteries will ever offer a positive expected rate of return.

8.1 Mega Millions / Powerball Note that for these lotteries we know the exact

break-even curves, so we needn’t use the general bound L Using the data from

Table 3B, we find that the break-even curve for Mega Millions is given by

8.2 Why Mega Millions and Powerball Will Always Be Bad Bets As we have

mentioned already, every time a Mega Millions or Powerball jackpot reaches J0,

Trang 21

the ticket sales N have easily exceeded the jackpot J Assuming this trend will

con-tinue, the preceding example shows that a gambler seeking a positive rate of return on

a Mega Millions or Powerball drawing need only consider those drawings where the

jackpot J is at least 2 J0 (Even then, of course, the drawing is not guaranteed to offer

a positive rate of return.) We give a heuristic calculation to show that the jackpot willprobably never be so large

Since the inception of these two games, the value of J /J0has exceeded 1 only threetimes, in the last three drawings in Table 4 The maximum value attained so far is 1.19

What would it take for J /J0to reach 2?

We will estimate two things: first, the probability that a large jackpot (J ≥ J0) rolls

over, and second, the number of times this has to happen for the jackpot to reach 2J0.The chance of a rollover is the chance that the jackpot is not won, i.e.,



1− 1

t

N

where N is the number of tickets sold This is a decreasing function of N , so in order

to find an upper bound for this chance, we need a lower bound on N as a function of J

For this we use the assumption (1) of Section 8.1, which historically has been satisfiedfor all large jackpots.eSince a rollover can only cause the jackpot to increase, the same

lower bound on N will hold for all future drawings until the jackpot is won Thus if the current jackpot is J ≥ J0, the chance of rolling over k times is at most



1− 1

t

k J0.

Now, J0 = Ft and since t is quite large, we may approximate (1 − 1/t) tby its limiting

value as t tends to infinity, which is 1 /e Thus the probability above is roughly e −k F.

Since F ≥ 0.82, we conclude that once the jackpot reaches J0, the chance of it rolling

over k times is no more than e −.82k

Let us now compute the number k of rollovers it will take for the jackpot to reach 2J0 Each time the jackpot (J ) rolls over (due to not being won), the jackpot for the next drawing increases, say to J The amount of increase depends on the ticket sales,

as this is the only source of revenue and a certain fraction of revenue is mandated to gotoward the jackpot So in order to predict how much the jackpot will increase, we need

a model of ticket sales as a function of the jackpot It is not clear how best to model this,

since the data are so sparse in this range For smaller jackpots, evidence [7] indicates

that ticket sales grow as a quadratic function of the jackpot, so one possibility is toextrapolate to this range On the other hand, no model can work for arbitrarily largejackpots, because of course at some point the sales market will be saturated For the

purposes of this argument, then, we note that the ratio J /J has never been larger than

about 1.27 for any reasonably large jackpot, and we use this as an upper bound on the

rolled-over jackpot This implicitly assumes a linear upper bound on ticket sales as afunction of the jackpot, but the bound is nevertheless generous by historical standards.f

So, suppose that each time a large jackpot J is rolled over, the new jackpot J isless than 1.27J Then, even the jackpot corresponding to the largest value of J/J0

e Added in proof: The 2008–2009 recession appears to have disturbed this trend The 8/28/09 Mega Millions

drawing had an announced jackpot of $335m, leading to J ≈ 159m About 149m tickets were sold, so N ≈

0.94J But this drawing was still a bad bet with an eRoR of −23%.

fIt was pointed out to us by Victor Matheson that over time lottery ticket sales are trending downward [7],

so our upper bound on all past events is likely to remain valid in the future.

Trang 22

(namely J /J0 = 1.19) would have had to roll over three more times before J would have surpassed 2 J0 Thus we should evaluate e −0.82k with k= 3.

Plugging in k = 3 shows that once the jackpot reaches J0, the probability that it

reaches 2J0is at most e −0.82·3 < 1/11 So if the jackpot of one of these lotteries reaches

J0about once every two years (a reasonable estimate by historical standards), then one

would expect it to reach 2J0 about once every 22 years, which is longer than the lifespan of most lotteries This is the basis of our claim that Mega Millions and Powerballare unlikely to ever offer a positive eRoR

As a final remark, we point out that even in the “once every 22 years” case where

the jackpot exceeds 2 J0, one only concludes that the hypotheses of Example 8.1 arenot satisfied; one still needs to check whether the expected rate of return is positive.And in any case, there still remains the question of whether buying lottery tickets is agood investment

P ART III TO INVEST OR NOT TO INVEST: ANALYZING THE RISK In the

last few sections we have seen that good bets in the lottery do exist, though they may

be rare We also have an idea of how to find them So if you are on the lookout andyou spot a good bet, is it time to buy tickets? In the rest of the paper we address thequestion of whether such an opportunity would actually make a good investment

9 Is Positive Rate of Return Enough? The Lotto Texas drawing of April 7, 2007

had a huge rate of return of 30% over a very short time period (which for the sake of gument we call a week) Should you buy tickets for such a drawing? The naive investorwould see the high rate of return and immediately say “yes.”gBut there are many riskyinvestments with high expected rates of return that investors typically choose not to in-vest in directly, like oil exploration or building a time machine How should one makethese decisions?

ar-This is an extreme example of a familiar problem: you want to invest in a nation of various assets with different rates of return and different variances in thoserates of return Standard examples of such assets are a certificate of deposit with a low(positive) rate of return and zero variance, bonds with a medium rate of return andsmall variance, or stocks with a high rate of return and large variance How can youdecide how much to invest in each? Undergraduate economics courses teach a method,

combi-known as mean-variance portfolio analysis, by which the investor can choose the

opti-mal proportion to invest in each of these assets (Harry Markowitz and William Sharpeshared the Nobel Prize in economics in 1990 for their seminal work in this area.) Wewill apply this same method to compare lottery tickets with more traditional risky as-sets This will provide both concrete investment advice and a good illustration of how

to apply the method, which requires little more than basic linear algebra

The method is based on certain assumptions about your investment preferences

You want to pick a portfolio, i.e., a weighted combination of the risky assets where the

weights sum to 1 We assume:

(1) When you decide whether or not to invest in an asset, the only attributes you sider are its expected rate of return, the variance in that rate, and the covariance

con-of that rate with the rates con-of return con-of other assets.h

g The naive view is that rate of return is all that matters This is not just a straw man; the writers of tual fund prospectuses often capitalize on this naive view by reporting historical rates of return, but not, say, correlation with the S&P 500.

mu-h This assumption is appealing from the point of view of mathematical modeling, but there are various objections to it, hence also to mean-variance analysis But we ignore these concerns because mean-variance

analysis is “the de-facto standard in the finance profession” [5,§2.1].

Trang 23

(2) You prefer portfolios with high rate of return and low variance over portfolioswith low rate of return and high variance (This is obvious.)

(3) Given a choice between two portfolios with the same variance, you prefer theone with the higher rate of return

(4) Given a choice between two portfolios with the same rate of return, you preferthe one with the lower variance

From these assumptions, we can give a concrete decision procedure for picking anoptimal portfolio

But before we do that, we first justify (4) In giving talks about this paper, we havefound that (4) comes as a surprise to some mathematicians; they do not believe that itholds for every rational investor Indeed, if an investor acts to maximize her expectedamount of money, then two portfolios with the same rate of return would be equallydesirable But economists assume that rational investors act to maximize their expected

utility.i

In mathematical terms, an investor derives utility (pleasure) U (x) from having x dollars, and the investor seeks to maximize the expected value of U (x) Economists assume that U (x) is positive, i.e., the investor always wants more money They call this axiom non-satiety They also assume that U (x) is negative, meaning that some-

one who is penniless values a hundred dollars more than a billionaire does These sumptions imply (4), as can be seen by Jensen’s inequality: given two portfolios withthe same rate of return, the one with lower variance will have higher expected utility

as-As an alternative to using Jensen’s inequality, one can examine the Taylor polynomial

of degree 2 for U (x) The great economist Alfred Marshall took this second approach

in Note IX of the Mathematical Appendix to [19].j Economists summarize property

(4) by saying that rational investors are risk-averse.

10 Example of Portfolio Analysis We consider a portfolio consisting of typical

risky assets: the iShares Barclays aggregate bond fund (AGG), the MSCI EAFE dex (which indexes stocks from outside of North America), the FTSE/NAREIT allREIT index (indexing U.S Real Estate Investment Trusts), the Standard & Poors 500(S&P500), and the NASDAQ Composite index We collected weekly adjusted returns

in-on these investments for the period January 31, 1972 through June 4, 2007—exceptfor AGG, for which the data began on September 29, 2003 This time period does notinclude the current recession that began in December 2007 The average weekly rates

of return and the covariances of these rates of return are given in Table 10

Table 10 Typical risky investments The third column gives the expected weekly rate of return in % Columns

4–8 give the covariances (in % 2 ) between the weekly rates of return The omitted covariances can be filled in

eval-the utility eval-they can obtain from it.” [3, p 33].

j That note ends with: “ experience shows that [the pleasures of gambling] are likely to engender a

restless, feverish character, unsuited for steady work as well as for the higher and more solid pleasures of life.”

Trang 24

We choose one of the simplest possible versions of mean-variance analysis Weassume that you can invest in a risk-free asset—e.g., a short-term government bond or

a savings account—with positive rate of return R F The famous “separation theorem”says that you will invest in some combination of the risk-free asset and an “efficient”

portfolio of risky assets, where the risky portfolio depends only on R F and not on your

particular utility function See [29] or any book on portfolio theory for more on this

theorem

We suppose that you will invest an amount i in the risky portfolio, with units chosen

so that i = 1 means you will invest the price of 1 lottery ticket We describe the risky

portfolio with a vector X such that your portfolio contains i X k units of asset k If X k

is negative, this means that you “short” i |X k | units of asset k We require further that



k |X k| = 1, i.e., in order to sell an asset short, you must put up an equal amount

of cash as collateral (An economist would say that we allow only “Lintnerian” short

sales, cf [16].)

Determining X is now an undergraduate exercise We follow §II of [16] Writeμ

for the vector of expected returns, so thatμ k is the eRoR on asset k, and write C for

the symmetric matrix of covariances in rates of return; initially we only consider the 5

typical risky assets from Table 10 Then by [16, p 21, (14)],

11 Should You Invest in the Lottery? We repeat the computation from the previous

section, but we now include a particular lottery drawing as asset 6 We suppose that

the drawing has eRoR R L and variancev Recall that the eRoR is given by (4.5) One definition of the variance is E (X2) − E(X)2; here is a way to quickly estimate it

11.1 Estimating the Variance For lottery drawings, E (X2) dwarfs E(X)2 and byfar the largest contribution to the former term comes from the jackpot Recall thatjackpot winnings might be shared, so by taking into account the odds for the variouspossible payouts, we get the following estimate for the variancev1on the rate of returnfor a single ticket, where w denotes the total number of tickets winning the jackpot,

There is an extra complication The variance of an investment in the lottery depends

on how many tickets you buy, and we avoid this worry by supposing that you are

buy-ing shares in a syndicate that expects to purchase a fixed number S of lottery tickets.

Trang 25

In this way, the variancev of your investment in the lottery is the same as the variance

in the syndicate’s investment Assuming the number S of tickets purchased is small

relative to the total number of possible tickets, the variancev of your investment is

approximatelyv1/S.

11.2 The Lottery as a Risky Asset With estimates of the eRoR R Land the variance

v in hand, we can consider the lottery as a risky asset, like stocks and bonds We write

ˆμ for the vector of expected rates of return and ˆC for the matrix of covariances, so that

amount of “nearly all” of the possible securities In the real world, finance als do not invest in assets where mean-variance analysis suggests an investment of lessthan some fractionθ of the total Let us do this With that in mind, should you invest

profession-in the lottery?

Negative Theorem Under the hypotheses of the preceding paragraph, if

v ≥ R L − R F

0.022θ ,

then an efficient portfolio contains a negligible fraction of lottery tickets.

Before we prove this Negative Theorem, we apply it to the Lotto Texas drawing

from Table 4, so, e.g., R Lis about 30% and the variance for a single ticketv1is about

4× 1011(see Section 11.1) Suppose that you have $1000 to invest for a week in somecombination of the 5 typical risky assets or in tickets for such a lottery drawing Wewill round our investments to whole numbers of dollars, so a single investment of lessthan 50 cents will be considered negligible Thus we takeθ = 1/2000 We have:

R L − R F 022θ <

30

1.1 × 10−5 < 2.73 × 106.

In order to see if the Negative Theorem applies, we want to know whetherv is

bigger than 2.73 million As in Section 11.1, we suppose that you are buying shares in a syndicate that will purchase S tickets, so the variance v is approximately (4 × 1011)/S.

That is, it appears that the syndicate would have to buy around

S≈ 4× 1011

2.73 × 106 ≈ 145,000 tickets

to make the (rather coarse) bound in the Negative Theorem fail to hold

Trang 26

In other words, according to the Negative Theorem, if the syndicate buys fewerthan 145,000 tickets, our example portfolio would contain zero lottery tickets (If the

syndicate buys more than that number of tickets, then it might be worth your while

to get involved.) This is the quantitative statement we were seeking to validate ourintuition: buying lottery tickets really is a bad investment

Proof of the Negative Theorem We want to prove that | ˆX6| < θ We have:

ˆZ2 > 0.022 Plugging in the formula | ˆZ6| = (R L − R F )/v, we find:

| ˆX6| < R L − R F

0.022v .

By hypothesis, the fraction on the right is at mostθ.

12 Conclusions So, how can you make money at the lottery?

Positive return If you just want a positive expected rate of return, then our results in

Section 8 say to avoid Mega Millions and Powerball Instead, you should buy tickets

in lotteries where the ticket sales are a small fraction of the jackpot And furthermore,you should only do so when the jackpot is unusually large—certainly the jackpot has

to be bigger than the jackpot cutoff J0defined in (3.3)

To underline these results, we point out that the Small Ticket Sales example 7.1gives concrete criteria that guarantee that a lottery drawing has a positive rate of re-turn We remark that even though these drawings have high variances, a positive rate

of return is much more desirable than, say, casino games like roulette, keno, and slotmachines, which despite their consistent negative rates of return are nonetheless ex-tremely popular

Positive return and buyouts For drawings with a positive rate of return, one can

obtain a still higher rate of return by “buying out” the lottery, i.e., buying one of each

of the possible distinct tickets We omit a detailed analysis here because of variouscomplicating factors such as the tax advantages, increase in jackpot size due to extraticket sales, and a more complicated computation of expected rate of return due to thelower-tier pari-mutuel prizes The simple computations we have done do not scale up

to include the case of a buyout See [11] for an exhaustive empirical study.

In any case, buying out the lottery requires a large investment—in the case of LottoTexas, about $26 million This has been attempted on several occasions, notably with

Australia’s New South Wales lottery in 1986, the Virginia (USA) Lottery in 1992 [22], and the Irish National Lottery in 1992 [10].kBuying out the lottery may be tax advan-taged because the full cost of buying the tickets may be deductible from the winnings;please consult a tax professional for advice On the other hand, buying out the lotteryincurs substantial operational overhead in organizing the purchase of so many tickets,which typically must be purchased by physically going to a lottery retailer and fillingout a play slip

k Interestingly, in both the Virginia and Irish drawings, buyout organizers won the jackpot despite only

buying about 70% [33] and 92% [8] of the possible tickets, respectively.

Trang 27

We also point out that in some of the known buyout attempts the lottery companiesresisted paying out the prizes, claiming that the practice of buying out the game is

counter to the spirit of the game (see [22], [10]) In both the cited cases there was

a settlement, but one should be aware that a buyout strategy may come up againstsubstantial legal costs

Investing If you are seeking good investment opportunities, then our results in

Sec-tion 11 suggest that this doesn’t happen in the lottery, due to the astounding variances

in rates of return on the tickets

What if a syndicate intends to buy out a lottery drawing, and there is a positiveexpected rate of return? Our mean-variance analysis suggests that you should invest asmall amount of money in such a syndicate

Alternative strategies Buying tickets is not the only way to try to make money off

the lottery The corporations that operate Mega Millions and Powerball are both itable, and that’s even after giving a large fraction of their gross income (35% in the

prof-case of Mega Millions [25]) as skim to the participating states Historically ing, lotteries used to offer a better rate of return to players [30], possibly because the

speak-lottery only had to provide profits to the operators With this in mind, today’s trepreneur could simply run his or her own lottery with the same profit margins butreturning to the players (in the form of better prizes) the money that the state wouldnormally take This would net a profit for the operator and give the lottery players abetter game to play Unfortunately, we guess that running one’s own lottery is illegal

en-in most cases But a similar option may be open to casen-inos: a small modification ofkeno to allow rolling jackpots could combine the convenience and familiarity of kenowith the excitement and advertising power of large jackpots

ACKNOWLEDGMENTS We thank Amit Goyal and Benji Shemmer for sharing their helpful insights; Bill

Ziemba and Victor Matheson for valuable feedback; and Rob O’Reilly for help with finding the data in Table

10 We owe Ron Gould a beer for instigating this project The second author was partially supported by NSF grant DMS-0653502.

REFERENCES

1. D R Bellhouse, The Genoese lottery, Statist Sci 6 (1991) 141–148.doi:10.1214/ss/1177011819

2. , Euler and lotteries, in Leonhard Euler: Life, Work and Legacy, R E Bradley, C E Sandifer,

and E Sandifer, eds., Elsevier, Amsterdam, 2007.

3. D Bernoulli, Specimen theoriae novae de mensura sortis, Commentarii Academiae Scientiarum

Imperi-alis Petropolitanae V (1738) 175-192; English translation by L Sommer, Exposition of a new theory on

the measurement of risk, Econometrica 22 (1954) 23–36.

4 C W Blau and J B Coutant, Federal tax treatment of gains and losses from gambling transactions,

Gaming Law Review 7 (2003) 319–322.doi:10.1089/109218803770238434

5. M W Brandt, Portfolio choice problems, in Handbook of Financial Econometrics, L Hansen and Y

Ait-Sahalia, eds., North Holland, Amsterdam, 2009.

6. H Chernoff, How to beat the Massachussetts numbers game, Math Intelligencer 3 (1981) 166–172.

9. L Euler, Sur la probabilit´e des s´equences dans la lotterie G´enoise, Memoires de l’academie des

sci-ences de Berlin 21 (1767) 191–230; also available athttp://math.dartmouth.edu/~euler/docs/ originals/E338.pdf.

10. R Fowler, How to make a killing on the lottery, The Independent (London), January 4, 1996.

Trang 28

11. K R Grote and V A Matheson, In search of a fair bet in the lottery, Eastern Economic Journal 32

(2006) 673–684.

12. J Haigh, The statistics of lotteries, in Handbook of Sports and Lottery Markets, D B Hausch and W T.

Ziemba, eds., North Holland, Amsterdam, 2008.

13. N Henze and H Riedwyl, How to Win More: Strategies for Increasing a Lottery Win, A K Peters, Natick,

16 J Lintner, The valuation of risk assets and the selection of risky investments in stock portfolios and

capital budgets, The Review of Economics and Statistics 47 (1965) 13–37.doi:10.2307/1924119

17 L C MacLean, W T Ziemba, and G Blazenko, Growth versus security in dynamic investment analysis,

Management Science 38 (1992) 1562–1585.doi:10.1287/mnsc.38.11.1562

18. Malcolm X and A Haley, The Autobiography of Malcolm X, Grove Press, New York, 1965.

19. A Marshall, Principles of Economics, 8th ed., Macmillan, London, 1920; also available athttp://www econlib.org/library/Marshall/marP.html.

20. V A Matheson, When are state lotteries a good bet (revisited)? Eastern Economic Journal 27 (2001)

55–70.

21. V A Matheson and K Grothe, US lotto markets, in Handbook of Sports and Lottery Markets, D B.

Hausch and W T Ziemba, eds., North Holland, Amsterdam, 2008.

22. I McMinn, Number is up for Lotto whizz, Sunday Herald Sun (Melbourne), April 30, 1995.

23 D Nettles, Total sales and jackpot amounts for each of the Lotto Texas drawings, available at http: //www.lottoreport.com/lottosales.htm, retrieved August 10, 2007.

24. New Jersey Division of Taxation, Form NJ-1040 line-by-line instructions (2006), available athttp: //www.state.nj.us/treasury/taxation/pdf/other_forms/tgi-ee/2006/061040i.pdf.

25 Official home of Mega Millions, available at http://www.megamillions.com/aboutus/where_ money_goes.asp, retrieved February 19, 2008.

26. W Parry, Who’s laughing now? Jokester is $258 million richer, Associated Press wire, September 23,

2005.

27 Powerball home page, available at http://www.powerball.com, retrieved February 20, 2008.

28. L M Rotando and E O Thorp, The Kelly criterion and the stock market, American Mathematical

Monthly 99 (1992) 922–931.doi:10.2307/2324484

29. W F Sharpe, Portfolio Theory and Capital Markets, McGraw-Hill, New York, 1970.

30. R C Sprowls, A historical analysis of lottery terms, The Canadian Journal of Economics and Political

Science 20 (1954) 347–356.doi:10.2307/138511

31 The Straight Dope, If I win the lottery, should I ask for a lump sum or an annuity?, available at http: //www.straightdope.com/mailbag/mlottery.html, retrieved August 6, 2007.

32. E O Thorpe, The Kelly criterion in blackjack, sports betting, and the stock market, in Handbook of Asset

and Liability Management, vol 1: Theory and Methodology, S A Zenios and W T Ziemba, eds., North

Holland, Amsterdam, 2006.

33. The Washington Post, Va Lotto payoff approved, March 11, 1992.

34. W T Ziemba, S L Brumelle, A Gautier, and S L Schwartz, Dr Z’s 6/49 Lotto Guidebook, Dr Z

Investments, Vancouver, 1986.

35. W T Ziemba, Efficiency of racing, sports and lottery betting markets, in Handbook of Sports and Lottery

Markets, D B Hausch and W T Ziemba, eds., North Holland, Amsterdam, 2008.

AARON ABRAMS is assistant professor of mathematics at Emory University He studies mathematical

prob-lems in a variety of fields, including topology, group theory, sports, and, apparently, how to make money As a general rule, he does not buy lottery tickets.

Department of Mathematics & Computer Science, 400 Dowman Dr., Emory University, Atlanta, Georgia 30322 abrams@mathcs.emory.edu

SKIP GARIBALDI is Winship Distinguished Professor of mathematics at Emory University, where he also

teaches some mathematical economics His retirement savings make up a well-diversified portfolio, but since

he invested starting in 2000, he would have more money if he had just kept it all in his mattress.

Department of Mathematics & Computer Science, 400 Dowman Dr., Emory University, Atlanta, Georgia 30322 skip@member.ams.org

Trang 29

Voting in Agreeable Societies

Deborah E Berg, Serguei Norine, Francis Edward Su,

Robin Thomas, and Paul Wollan

My idea of an agreeable person is a person who agrees with me.

—Benjamin Disraeli [1, p 29]

1 INTRODUCTION When is agreement possible? An important aspect of group

decision-making is the question of how a group makes a choice when individual erences may differ Clearly, when making a single group choice, people cannot allhave their “ideal” preferences, i.e., the options that they most desire, if those idealpreferences are different However, for the sake of agreement, people may be willing

pref-to accept as a group choice an option that is merely “close” pref-to their ideal preferences.Voting is a situation in which people may behave in this way The usual startingmodel is a one-dimensional political spectrum, with conservative positions on the rightand liberal positions on the left, as in Figure 1 We call each position on the spectrum

a platform that a candidate or voter may choose to adopt While a voter may represent her ideal platform by some point x on this line, she might be willing to vote for a candidate who is positioned at some point “close enough” to x, i.e., in an interval about x.

x

[ I ]

Figure 1 A one-dimensional political spectrum, with a single voter’s interval of approved platforms.

In this article, we ask the following: given such preferences on a political spectrum,when can we guarantee that some fraction (say, a majority) of the population will

agree on some candidate? By “agree”, we mean in the sense of approval voting, in

which voters declare which candidates they find acceptable

Approval voting has not yet been adopted for political elections in the United States.However, many scientific and mathematical societies, such as the Mathematical Asso-ciation of America and the American Mathematical Society, use approval voting fortheir elections Additionally, countries other than the United States have used approval

voting or an equivalent system; for details, see Brams and Fishburn [3] who discuss

the advantages of approval voting

Understanding which candidates can get voter approval can be helpful when thereare a large number of candidates An extreme example is the 2003 California guber-

natorial recall election, which had 135 candidates in the mix [7] We might imagine

these candidates positioned at 135 points on the line in Figure 1, which we think of as

a subset ofR If each California voter approves of candidates “close enough” to herideal platform, we may ask under what conditions there is a candidate that wins theapproval of a majority of the voters

doi:10.4169/000298910X474961

Trang 30

In this setting, we may assume that each voter’s set of approved platforms (her proval set) is a closed interval in R, and that there is a set of candidates who take

ap-up positions at various points along this political spectrum We shall call this trum with a collection of candidates and voters, together with voters’ approval sets,

spec-a linespec-ar society (spec-a more precise definition will be given soon) We shspec-all sspec-ay thspec-at the linear society is super-agreeable if for every pair of voters there is some candidate that

they would both approve, i.e., each pair of approval sets contains a candidate in theirintersection For linear societies this “local” condition guarantees a strong “global”

property, namely, that there is a candidate that every voter approves! As we shall see in

Theorem 5, this can be viewed as a consequence of Helly’s theorem about intersections

of convex sets

But perhaps this is too strong a conclusion Is there a weaker local condition thatwould guarantee that only a majority (or some other fraction) of the voters wouldapprove a particular candidate? For instance, we relax the condition above and call a

linear society agreeable if among every three voters, some pair of voters approve the

same candidate Then it is not hard to show:

Theorem 1 In an agreeable linear society, there is a candidate who has the approval

of at least half the voters.

More generally, call a linear society(k, m)-agreeable if it has at least m voters, and among every m voters, some subset of k voters approve the same candidate Then our

main theorem is a generalization of the previous result:

Theorem 2 (The Agreeable Linear Society Theorem) Let 2 ≤ k ≤ m In a (k, agreeable linear society of n voters, there is a candidate who has the approval of at least n (k − 1)/(m − 1) of the voters.

m)-We prove a slightly more general result in Theorem 8 and also briefly study societieswhose approval sets are convex subsets ofRd

As an example, consider a city with fourteen restaurants along its main boulevard:

A B C D E F G H I J K L M N

and suppose every resident dines only at the five restaurants closest to his/her house (a

set of consecutive restaurants, e.g., D E F G H ) A consequence of Theorem 1 is that

there must be a restaurant that is patronized by at least half the residents Why? Thepigeonhole principle guarantees that among every 3 residents, each choosing 5 of 14restaurants, there must be a restaurant approved by at least 2 of them; hence this linearsociety is agreeable and Theorem 1 applies For an example of Theorem 2, see Figure

3, which shows a(2, 4)-agreeable linear society, and indeed there are candidates that

receive at least 1/3 of the votes (in this case 7/3 = 3).

We shall begin with some definitions, and explain connections to classical convexitytheorems, graph colorings, and maximal cliques in graphs Then we prove Theorem 2,discuss extensions to higher-dimensional spectra, and conclude with some questionsfor further study

2 DEFINITIONS In this section, we fix terminology and explain the basic concepts

upon which our results rely Let us suppose that the set of all possible preferences is

modeled by a set X , called the spectrum Each element of the spectrum is a platform Assume that there is a finite set V of voters, and each voter v has an approval set A v

of platforms

Trang 31

We define a society S to be a triple (X, V, A ) consisting of a spectrum X, a set of voters V , and a collection Aof approval sets for all the voters Of particular interest to

us will be the case of a linear society, in which X is a closed subset ofR and approvalsets inA are of the form X ∩ I where I is either empty or a closed bounded interval

inR In general, however, X could be any set and the collection Aof approval sets

could be any class of subsets of X In Figure 2 we illustrate a linear society, where

for ease of display we have separated the approval sets vertically so that they can bedistinguished

1

2 3

4

6 5 7

Figure 2 A linear society with infinite spectrum: each interval (shown here displaced above the spectrum)

corresponds to the approval set of a voter The shaded region indicates platforms with agreement number 4 This is a(2, 3)-agreeable society.

Our motivation for considering intervals as approval sets arises from imagining

that voters have an “ideal” platform along a linear scale (similar to Coombs’ J -scale

[8]), and that voters are willing to approve “nearby” platforms, yielding approval sets

that are connected intervals Unlike the Coombs scaling theory, however, we are notconcerned with the order of preference of approved platforms; all platforms within avoter’s approval set have equivalent status as “approved” by that voter We also notethat while we model our linear scale as a subset ofR, none of our results about linearsocieties depends on the metric; we only appeal to the ordinal properties ofR

We have seen that politics provides natural examples of linear societies For a

dif-ferent example, X could represent a temperature scale, V a set of people that live in a house, and each A va range of temperatures that personv finds comfortable Then one

may ask: at what temperature should the thermostat be set so as to satisfy the largestnumber of people?

1

2 3

4

6 5 7

Figure 3 A linear society with a spectrum of two candidates (at platforms marked by carats): take the approval

sets of the society of Figure 2 and intersect with these candidates It is a(2, 4)-agreeable linear society.

Two special cases of linear societies are worth mentioning When X = R we should

think of X as an infinite spectrum of platforms that potential candidates might adopt.

However, in practice there are normally only finitely many candidates We model that

situation by letting X be the set of platforms adopted by actual candidates Thus one could think of X as either the set of all platforms, or the set of (platforms adopted by)

candidates See Figures 2 and 3

Trang 32

Let 1≤ k ≤ m be integers Call a society (k, m)-agreeable if it has at least m voters, and for any subset of m voters, there is at least one platform that at least k of them can agree upon, i.e., there is a point common to at least k of the voters’ approval sets Thus

to be(2, 3)-agreeable is the same as to be agreeable, and to be (2, 2)-agreeable is the same as to be super-agreeable, as defined earlier.

One may check that the society of Figure 2 is (2, 3)-agreeable It is not (3,

4)-agreeable, however, because among voters 1, 2, 4, 7 no three of them share a commonplatform The same society, after restricting the spectrum to a set of candidates, is thelinear society shown in Figure 3 It is not(2, 3)-agreeable, because among voters 2, 4,

7 there is no pair that can agree on a candidate (in fact, voter 7 does not approve anycandidate) However, one may verify that this linear society is(2, 4)-agreeable For a society S, the agreement number of a platform, a (p), is the number of voters

in S who approve of platform p The agreement number a (S) of a society S is the

maximum agreement number over all platforms in the spectrum, i.e.,

3 HELLY’S THEOREM AND SUPER-AGREEABLE SOCIETIES Let us say

that a society isRd -convex if the spectrum isRd and each approval set is a closed vex subset ofRd Note that anR1-convex society is a linear society with spectrumR

con-AnRd

-convex society can arise when considering a multi-dimensional spectrum, such

as when evaluating political platforms over several axes (e.g., conservative vs liberal,pacifist vs militant, interventionist vs isolationist) Or, the spectrum might be arrayed

over more personal dimensions: the dating website eHarmony claims to use up to 29 of

them [9] In such situations, the convexity of approval sets might, for instance, follow

from an independence-of-axes assumption and convexity of approval sets along eachaxis

To find the agreement proportion of anRd-convex society, we turn to work cerning intersections of convex sets The most well-known result in this area is Helly’stheorem This theorem was proven by Helly in 1913, but the result was not published

con-until 1921, by Radon [17].

Theorem 3 (Helly) Given n convex sets in Rd where n > d, if every d + 1 of them intersect at a common point, then they all intersect at a common point.

Helly’s theorem has a nice interpretation forRd-convex societies:

Corollary 4 For every d ≥ 1, a (d + 1, d + 1)-agreeable R d -convex society must contain at least one platform that is approved by all voters.

Notice that for the corollary to hold for d > 1 it is important that the spectrum of an

Rd -convex society be all ofRd However, for d= 1 that is not necessary, as we nowshow

Theorem 5 (The Super-Agreeable Linear Society Theorem) A super-agreeable

linear society must contain at least one platform that is approved by all voters.

Trang 33

We provide a simple proof of this theorem, since the result will be needed later.When the spectrum is all ofR, this theorem is just Helly’s theorem for d = 1; a proof

of Helly’s theorem for general d may be found in [15].

Proof Let X ⊆ R denote the spectrum Since each voter v agrees on at least one platform with every other voter, we see that the approval sets A vmust be nonempty Let

L v = min A v , R v = max A v , and let x= maxv {L v } and y = min v {R v} The first two

expressions exist because each A v is compact; the last two exist because the number

of voters is finite

^

y x

Figure 4 A super-agreeable linear society of 6 voters and 4 candidates, with agreement number 6.

We claim that x ≤ y Why? Since every pair of approval sets intersect in some platform, we see that L i ≤ R j for every pair of voters i , j In particular, let i be the voter whose L i is maximal and let j be the voter whose R j is minimal Hence x ≤ y and every approval set contains all platforms of X that are in the nonempty interval [x, y], and in particular, the platform x.

The idea of this proof can be easily extended to furnish a proof of Theorem 1

Proof of Theorem 1 Using the same notations as in the prior proof, if x ≤ y then that proof shows that every approval set contains the platform x Otherwise x > y implies

L i > R j so that A j and A ido not contain a common platform

We claim that for any other voterv, the approval set A v contains either platform

x or y (or both) After all, the society is agreeable, so some pair of A i , A j , A v must

contain a common platform; by the remarks above it must be that A vintersects one of

A i or A j If A v does not contain x = L i then since L v ≤ L i (by definition of x), we must have that R v < L i and A v ∩ A i does not contain a platform Then A v ∩ A jmust

contain a platform; hence L v ≤ R j Since R j ≤ R v (by definition of y), the platform

y = R j must be in A v

Thus every approval set contains either x or y, and by the pigeonhole principle one

of them must be contained in at least half the approval sets

Proving the more general Theorem 2 will take a little more work

4 THE AGREEMENT GRAPH OF LINEAR SOCIETIES IS PERFECT To

understand(k, m)-agreeability, it will be helpful to use a graph to represent the tersection relation on approval sets Recall that a graph G consists of a finite set V (G)

in-of vertices and a set E (G) of 2-element subsets of V (G), called edges If e = {u, v}

is an edge, then we say that u , v are the ends of e, and that u and v are adjacent in G.

We use u v as notation for the edge e.

Given a society S, we construct the agreement graph G of S by letting the vertices

V (G) be the voters of S and the edges E(G) be all pairs of voters u, v whose approval

Trang 34

sets intersect each other Thus u and v are connected by an edge if there is a platform that both u and v would approve Note that the agreement graph of a society with

agreement number equal to the number of voters is a complete graph (but the converse

is false in higher dimensions, as we discuss later) Also note that a vertexv is isolated

if A v is empty or disjoint from other approval sets

6 5

2 1

3

Figure 5 The agreement graph for the society in Figure 2 Note that voters 4, 5, 6, 7 form a maximal clique

that corresponds to the maximal agreement number in Figure 2.

The clique number of G, written ω(G), is the greatest integer q such that G has a set of q pairwise adjacent vertices, called a clique of size q By restricting our attention

to members of a clique, and applying the Super-Agreeable Linear Society Theorem,

we see that there is a platform that has the approval of every member of a clique, andhence:

Fact 1 For the agreement graph of a linear society, the clique number of the graph is

the agreement number of the society.

This fact does not necessarily hold if the society is not linear For instance, it is easy

to construct anR2-convex society with three voters such that every two voters agree on

a platform, but all three of them do not It does, however hold inRd for box societies,

to be discussed in Section 6

Now, to get a handle on the clique number, we shall make a connection between

the clique number and colorings of the agreement graph The chromatic number of

G, written χ(G), is the minimum number of colors necessary to color the vertices

of G such that no two adjacent vertices have the same color Thus two voters may

have the same color as long as they do not agree on a platform Note that in all cases,

χ(G) ≥ ω(G).

A graph G is called an interval graph if we can assign to every vertex x a closed interval or an empty set I x ⊆ R such that xy ∈ E(G) if and only if I x ∩ I y

have:

Fact 2 The agreement graph of a linear society is an interval graph.

To see that Fact 2 holds let the linear society be (X, V, A ), and let the voter proval sets be A v = X ∩ I v , where I vis a closed bounded interval or empty We may

ap-assume that each I v is minimal such that A v = X ∩ I v; then the collection{I v : v ∈ V }

provides an interval representation of the agreement graph, as desired

An induced subgraph of a graph G is a graph H such that V (H) ⊆ V (G) and the edges of H are the edges of G that have both ends in V (H) If every induced subgraph

H of a graph G satisfies χ(H) = ω(H), then G is called a perfect graph; see, e.g.,

[18] The following is a standard fact [20] about interval graphs:

Theorem 6 Interval graphs are perfect.

Trang 35

Proof Let G be an interval graph, and for v ∈ V (G), let I vbe the interval representing

the vertexv Since every induced subgraph of an interval graph is an interval graph,

it suffices to show that χ(G) = ω, where ω = ω(G) We proceed by induction on

|V (G)| The assertion holds for the null graph, and so we may assume that |V (G)| ≥ 1,

and that the statement holds for all smaller graphs Let us select a vertexv ∈ V (G) such that I v is empty or the right end of I v is as small as possible It follows that the

elements of N , the set of neighbors of v in V (G), are pairwise adjacent because their intervals must all contain the right end of I v, and hence |N| ≤ ω − 1 See Figure 6.

By the inductive hypothesis, the graph G \{v} obtained from G by deleting v can be

colored usingω colors, and since v has at most ω − 1 neighbors, this coloring can be extended to a coloring of G, as desired.

I

I

v

w

Figure 6 If I v , I w intersect and the right end of I v is smaller than the right end of I w , then I wmust contain

the right end of I v.

The perfect graph property will allow us, in the next section, to make a crucial nection between the (k, m)-agreeability condition and the agreement number of the

con-society Given its importance in our setting, it is worth making a few comments abouthow perfect graphs appear in other contexts in mathematics, theoretical computer

science, and operations research The concept was introduced in 1961 by Berge [2],

who was motivated by a question in communication theory, specifically, the

determi-nation of the Shannon capacity of a graph [19] Chv´atal later discovered that a certain

class of linear programs always have an integral solution if and only if the

correspond-ing matrix arises from a perfect graph in a specified way [6, 18, 5] As pointed out

in [18], algorithms to solve semi-definite programs grew out of the theory of perfect graphs It has been proven recently [4] that a graph is perfect if and only if it has no

induced subgraph isomorphic to a cycle of odd length at least five, or a complement ofsuch a cycle

5 (k, m)-AGREEABLE LINEAR SOCIETIES We now use the connection

be-tween perfect graphs, the clique number, and the chromatic number to obtain a lowerbound for the agreement number of a(k, m)-agreeable linear society (Theorem 8) We

first need a lemma that says that in the corresponding agreement graph, the(k,

m)-agreeable condition prevents any coloring of the graph from having too many vertices

of the same color Thus, there must be many colors and, since the graph is perfect, theclique number must be large as well

Lemma 7 Given integers m ≥ k ≥ 2, let positive integers q, ρ be defined by the sion with remainder: m − 1 = (k − 1)q + ρ, where 0 ≤ ρ ≤ k − 2 Let G be a graph

divi-on n ≥ m vertices with chromatic number χ such that every subset of V (G) of size m includes a clique of size k Then n ≤ χq + ρ, or χ ≥ (n − ρ)/q.

Proof Let the graph be colored using the colors 1 , 2, , χ, and for i = 1, 2, , χ let C i be the set of vertices of G colored i We may assume, by permuting the col-

ors, that|C1| ≥ |C2| ≥ · · · ≥ |C χ | Since C1∪ C2∪ · · · ∪ C k−1is colored using k− 1

colors, it includes no clique of size k, and hence, |C1∪ C2∪ · · · ∪ C k−1| ≤ m − 1.

Trang 36

It follows that |C k−1| ≤ q, for otherwise |C1∪ C2∪ · · · ∪ C k−1| ≥ (k − 1)(q + 1) ≥ (k − 1)q + ρ + 1 = m, a contradiction Thus |C i | ≤ q for each i ≥ k and



,

and this bound is best possible It follows that this is also a lower bound on the ment number, and hence every linear (k, m)-agreeable society has agreement propor- tion at least (k − 1)/(m − 1).

agree-Proof By Fact 2 and Theorem 6 the graph G is perfect Thus the chromatic number

of G is equal to ω(G), and hence ω(G) ≥ (n − ρ)/q by Lemma 7, as desired.

The second assertion follows from Fact 1 and noting that(n − ρ)(m − 1) = n(k −

1)q + nρ − ρ(m − 1) = n(k − 1)q + ρ(n − m + 1) ≥ n(k − 1)q, from which we see

that(n − ρ)/q ≥ n(k − 1)/(m − 1).

Let us observe that the bound(n − ρ)/q in Theorem 8 is best possible Indeed, let I1, I2, , I q be disjoint intervals, for i = q + 1, q + 2, , n − ρ let I i = I i −q,

and let I n −ρ+1 , I n −ρ+2 , , I n be pairwise disjoint and disjoint from all the previous

intervals, e.g., see Figure 7 Then the society with approval sets I1, I2, , I nis(k,

m)-agreeable and its agreement graph has clique number(n − ρ)/q.

Figure 7 A linear(4, 15)-society with n = 21 voters Here q = 4 and ρ = 2, so the clique number is at least

(n − ρ)/q = 5.

The Agreeable Linear Society Theorem (Theorem 2) now follows as a corollary ofTheorem 8

6. Rd -CONVEX AND d-BOX SOCIETIES In this section we prove a

higher-dimensional analogue of Theorem 8 by giving a lower bound on the agreementproportion of a(k, m)-agreeable R d-convex society We need a different method than

our method for d = 1, because for d ≥ 2, neither Fact 1 nor Fact 2 holds.

Also, we remark that, unlike our results on linear societies, our results in this sectionabout the agreement proportion for platforms will not necessarily hold when restrictingthe spectrum to a finite set of candidates inRd

We shall use the following generalization of Helly’s theorem, due to Kalai [11].

Trang 37

Theorem 9 (The Fractional Helly’s Theorem) Let d ≥ 1 and n ≥ d + 1 be gers, let α ∈ [0, 1] be a real number, and let β = 1 − (1 − α)1/(d+1) Let F

inte-1, F2, , F n

be convex sets inRd and assume that for at least α n

d+1



of the (d + 1)-element index

in at least βn of the sets F1, F2, , F n

The following is the promised analogue of Theorem 8

Theorem 10 Let d ≥ 1, k ≥ 2, and m ≥ k be integers, where m > d Then every (k, m)-agreeable R d -convex society has agreement proportion at least

Proof Let S be a (k, m)-agreeable R d -convex society, and let A1, A2, , A n be

its voter approval sets Let us call a set I ⊆ {1, 2, , n} good if |I | = d + 1 and

good sets We will do this by counting in two different ways the number N of all

pairs(I, J), where I ⊆ J ⊆ {1, 2, , n}, I is good, and |J| = m Let g be the ber of good sets Since every good set is of size d + 1 and extends to an m-element

every m-element set J ⊆ {1, 2, , n} includes at least one k-element set K with

For d = 1, Theorem 10 gives a worse bound than Theorem 8, and hence for d ≥ 2,

the bound is most likely not best possible However, a possible improvement must use

a different method, because the bound in Theorem 9 is best possible.

A box in Rd is the Cartesian product of d closed intervals, and we say that a ciety is a d-box society if each of its approval sets is a box in Rd By projection

so-onto each axis, it follows from Theorem 5 that d-box societies satisfy the conclusion

of Fact 1 (namely, that the clique number equals the agreement number), and hencetheir agreement graphs capture all the essential information about the society Unfor-

tunately, agreement graphs of d-box societies are, in general, not perfect For instance,

there is a 2-box society (Figure 8) whose agreement graph is the cycle on five vertices;hence its chromatic number is 3 but its clique number is 2

For k ≤ m ≤ 2k − 2, the following theorem will resolve the agreement

propor-tion problem for all(k, m)-agreeable societies satisfying the conclusion of Fact 1, and

hence for all(k, m)-agreeable d-box societies where d ≥ 1 (Theorem 13).

Theorem 11 Let m , k ≥ 2 be integers with k ≤ m ≤ 2k − 2, and let G be a graph on

n ≥ m vertices such that every subset of V (G) of size m includes a clique of size k Then ω(G) ≥ n − m + k.

Before we embark on a proof let us make a few comments First of all, the bound

n − m + k is best possible, as shown by the graph consisting of a clique of size

n − m + k and m − k isolated vertices Second, the conclusion ω(G) ≥ n − m + k implies that every subset of V (G) of size m includes a clique of size k, and so the two statements are equivalent under the hypothesis that k ≤ m ≤ 2k − 2 Finally, this hy- pothesis is necessary, because if m ≥ 2k − 1, then for n ≥ 2(m − k + 1), the disjoint

Trang 38

Figure 8 A 2-box society whose agreement graph is a 5-cycle.

union of cliques of sizes

not its conclusion

A vertex cover of a graph G is a set Z ⊆ V (G) such that every edge of G has at least one end in Z We say a set S ⊆ V (G) is stable if no edge of G has both ends in

S We deduce Theorem 11 from the following lemma.

Lemma 12 Let G be a graph with vertex cover of size z and none of size z − 1 such that G \{v} has a vertex cover of size at most z − 1 for all v ∈ V (G) Then |V (G)| ≤ 2z.

Proof Let Z be a vertex cover of G of size z For every v ∈ V (G) − Z let Z v be a

vertex cover of G \{v} of size z − 1, and let X v = Z − Z v Then X vis a stable set For

X ⊆ Z let N(X) denote the set of neighbors of X outside Z We have v ∈ N(X v ) and

N (X v ) − {v} ⊆ Z v − Z, and so

|X v | = |Z − Z v | = |Z| − |Z ∩ Z v | = |Z v | + 1 − |Z ∩ Z v | = |Z v − Z| + 1

≥ |N(X v )|.

On the other hand, if X ⊆ Z is stable, then |N(X)| ≥ |X|, for otherwise (Z − X) ∪

N (X) is a vertex cover of G of size at most z − 1, a contradiction We have

|Z| ≥ X

vN (X

where both unions are over allv ∈ V (G) − Z, and hence |V (G)| ≤ 2z, as required.

To see that the second inequality holds let u , v ∈ V (G) − Z Then

|X u ∪ X v | = |X u | + |X v | − |X u ∩ X v | ≥ |N(X u )| + |N(X v )| − |N(X u ∩ X v )|

≥ |N(X u )| + |N(X v )| − |N(X u ) ∩ N(X v )| = |N(X u ) ∪ N(X v )|,

and, in general, the second inequality of (1) follows by induction on|V (G) − Z| Proof of Theorem 11 We proceed by induction on n If n = m, then the conclusion certainly holds, and so we may assume that n ≥ m + 1 and that the theorem holds for graphs on fewer than n vertices We may assume that m > k, for otherwise the

Trang 39

hypothesis implies that G is the complete graph We may also assume that G has two nonadjacent vertices, say x and y, for otherwise the conclusion holds Then in G, every clique contains at most one of x , y, so in the graph G\{x, y} every set of vertices

of size m − 2 includes a clique of size k − 1 Since k − 1 ≤ m − 2 ≤ 2(k − 1) − 2

we deduce by the inductive hypothesis that ω(G) ≥ ω(G\{x, y}) ≥ n − 2 − (m −

2) + k − 1 = n − m + k − 1 We may assume in the last statement that equality holds throughout, because otherwise G satisfies the conclusion of the theorem Let ¯ G denote the complement of G; that is, the graph with vertex set V (G) and edge set consisting

of precisely those pairs of distinct vertices of G that are not adjacent in G Let us notice that a set Q is a clique in G if and only if V (G) − Q is a vertex cover in ¯G.

Thus ¯G has a vertex cover of size m − k + 1 and none of size m − k Let t be the least integer such that t ≥ m and ¯G has an induced subgraph H on t vertices with no vertex cover of size m − k We claim that t = m Indeed, if t > m, then the minimality of t implies that H \{v} has a vertex cover of size at most m − k for every v ∈ V (H) Thus

by Lemma 12, t = |V (H)| ≤ 2(m − k + 1) ≤ m < t, a contradiction Thus t = m.

By hypothesis, the graph ¯H has a clique Q of size k, but V (H) − Q is a vertex cover

of H of size m − k, a contradiction.

Theorem 13 Let d ≥ 1 and m, k ≥ 2 be integers with k ≤ m ≤ 2k − 2, and let S be

a (k, m)-agreeable d-box society with n voters Then the agreement number of S is at least n − m + k, and this bound is best possible.

Proof The agreement graph G of S satisfies the hypothesis of Theorem 11, and hence

it has a clique of size at least n − m + k by that theorem Since d-box societies satisfy

the conclusion of Fact 1, the first assertion follows The bound is best possible, because

the graph consisting of a clique of size n − m + k and m − k isolated vertices is an

interval graph

7 DISCUSSION As we have seen, set intersection theorems can provide a useful

framework to model and understand the relationships between sets of preferences invoting, and this context leads to new mathematical questions We suggest several di-rections which the reader may wish to explore

Recent results in discrete geometry have social interpretations The piercing number

[12] of approval sets can be interpreted as the minimum number of platforms that are

necessary such that everyone has some platform of which he or she approves Set

intersection theorems on other spaces (such as trees and cycles) are derived in [16]

as an extension of both Helly’s theorem and the KKM lemma [13]; as an application

the authors show that in a super-agreeable society with a circular political spectrum,there must be a platform that has the approval of a strict majority of voters (in contrast

with Theorem 5) Chris Hardin [10] has recently provided a generalization to(k,

m)-agreeable societies with a circular political spectrum

What results can be obtained for other spectra? The most natural problem seems to

be to determine the agreement proportion forRd -convex and d-box (k, m)-agreeable societies The smallest case where we do not know the answer is d = 2, k = 2, and m = 3 Rajneesh Hegde (private communication) found an example of a (2, 3)-

agreeable 2-box society with agreement proportion 3/8 There may very well be a nice formula, because for every fixed integer d the agreement number of a d-box so-

ciety can be computed in polynomial time This is because the clique number of the

corresponding agreement graph (also known as a graph of boxicity at most d) can be determined by an easy polynomial-time algorithm On the other hand, for every d ≥ 2

it is NP-hard to decide whether an input graph has boxicity at most d [14, 21] (For

Trang 40

d = 1 this is the same as testing whether a graph is an interval graph, and that can bedone efficiently.)

Passing from results about platforms in societies to results about a finite set of didates appears to be tricky in dimensions greater than 1 Are there techniques oradditional hypotheses that would give useful results about the existence of candidateswho have strong approval in societies with multi-dimensional spectra?

can-We may also question our assumptions While convexity seems to be a rationalassumption for approval sets in the linear case, in multiple dimensions additional con-siderations may become important One might also explore the possibility of discon-nected approval sets: what is the agreement proportion of a(k, m)-agreeable society

in which every approval set has at most two components?

One might also consider varying levels of agreement For instance, in a d-box

so-ciety, two voters might not agree on every axis, so their approval sets do not intersect,but it might be the case that many of the projections of their approval sets do In thiscase, one may wish to consider an agreement graph with weighted edges

Finally, we might wonder about the agreement parameters k and m for various

real-world issues For instance, a society considering outlawing murder would probably bemuch more agreeable than that same society considering tax reform Currently, we canempirically measure these parameters only by surveying large numbers of people about

their preferences It is interesting to speculate about methods for estimating suitable k and m from limited data.

This article grew out of the observation that Helly’s theorem, a classical result inconvex geometry, has an interesting voting interpretation This led to the development

of mathematical questions and theorems whose interpretations yield desirable sions in the voting context, e.g., Theorems 2, 8, 10, and 13 It is nice to see that whenclassical theorems have interesting social interpretations, the social context can alsomotivate the study of new mathematical questions

conclu-ACKNOWLEDGMENTS The authors wish to thank Steven Brams for helpful feedback and a referee for

many valuable suggestions The authors also gratefully acknowledge support by these NSF Grants: Berg

by 0301129, Norine by 0200595, Su by 0301129 and 0701308, Thomas by

DMS-0200595 and DMS-0354742.

REFERENCES

1. R Andrews, ed., The Columbia Dictionary of Quotations, Columbia University Press, New York, 1993.

2 C Berge, F¨arbung von Graphen deren s¨amtliche bzw deren ungerade Kreise starr sind

(Zusammen-fassung), Wissenschaftliche Zeitschrift, Martin Luther Universit¨at Halle-Wittenberg,

Mathematisch-Naturwissenschaftliche Reihe (1961) 114–115.

3 S J Brams and P C Fishburn, Going from theory to practice: The mixed success of approval voting,

Social Choice and Welfare 25 (2005) 457–474.doi:10.1007/s00355-005-0013-y

4. M Chudnovsky, N Robertson, P Seymour, and R Thomas, The strong perfect graph theorem, Ann of

Math 164 (2006) 51–229.doi:10.4007/annals.2006.164.51

5. , Progress on perfect graphs, Math Program Ser B 97 (2003) 405–422.

6. V Chv´atal On certain polytopes associated with graphs, J Combin Theory Ser B 13 (1975) 138–154.

doi:10.1016/0095-8956(75)90041-6

7 CNN.com, “California ballot official, ready to print,” August 14, 2003, available at http://www.cnn com/2003/ALLPOLITICS/08/13/calif.recall/index.html.

8. C H Coombs, A Theory of Data, John Wiley, New York, 1964.

9 eHarmony.com, “29 Dimensions TM of Compatibility” (2008), available at http://www.eharmony com/why/dimensions.

10. C S Hardin, Agreement in circular societies, Amer Math Monthly 117 (2010) 40–49.doi:10.4169/ 000298910X474970

... We may assume in the last statement that equality holds throughout, because otherwise G satisfies the conclusion of the theorem Let ¯ G denote the complement of G; that is, the graph with vertex... with n voters Then the agreement number of S is at least n − m + k, and this bound is best possible.

Proof The agreement graph G of S satisfies the hypothesis of Theorem 11, and... out of the observation that Helly’s theorem, a classical result inconvex geometry, has an interesting voting interpretation This led to the development

of mathematical questions and theorems

Ngày đăng: 25/05/2014, 22:12

Nguồn tham khảo

Tài liệu tham khảo Loại Chi tiết
1. M. Abramowitz and I. A. Stegun, eds., Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, Dover, New York, 1972; also available at http://dlmf.nist.gov/ Sách, tạp chí
Tiêu đề: Handbook of Mathematical Functions with Formulas, Graphs,"and Mathematical Tables
2. D. Borwein and J. M. Borwein, Some remarkable properties of sinc and related integrals, Ramanujan J.5 (2001) 73–90. doi:10.1023/A:1011497229317 Sách, tạp chí
Tiêu đề: Ramanujan J
3. D. Borwein, J. M. Borwein, and B. Mares, Multi-variable sinc integrals and volumes of polyhedra, Ra- manujan J. 6 (2002) 189–208. doi:10.1023/A:1015727317007 Sách, tạp chí
Tiêu đề: Ra-"manujan J
4. T. J. Bromwich, Theory of Infinite Series, 2nd ed., Blackie &amp; Sons, Glasgow, 1926 Sách, tạp chí
Tiêu đề: Theory of Infinite Series
5. H. S. Carslaw, An Introduction to the Theory of Fourier’s Series and Integrals, 3rd revised ed., Dover, New York, 1952 Sách, tạp chí
Tiêu đề: An Introduction to the Theory of Fourier’s Series and Integrals
6. N. G. de Bruijn, Asymptotic Methods in Analysis, 2nd ed., North-Holland, Amsterdam, 1961 Sách, tạp chí
Tiêu đề: Asymptotic Methods in Analysis
7. W. B. Gearhart and H. S. Schultz, The function sin( x x ) , College Math. J. 21 (1990) 90–99. doi:10.2307/2686748 Sách, tạp chí
Tiêu đề: x"x),College Math. J
8. P. Henrici, Applied and Computational Complex Analysis, Volume 2, Wiley, New York, 1977 Sách, tạp chí
Tiêu đề: Applied and Computational Complex Analysis
Tác giả: P. Henrici
Nhà XB: Wiley
Năm: 1977
9. I. E. Leonard and J. Duemmel, More—and Moore—power series without Taylor’s theorem, American Mathematical Monthly 92 (1985) 588–589. doi:10.2307/2323175 Sách, tạp chí
Tiêu đề: More—and Moore—power series without Taylor’s theorem
Tác giả: I. E. Leonard, J. Duemmel
Nhà XB: American Mathematical Monthly
Năm: 1985
10. F. W. J. Olver, Asymptotics and Special Functions, 2nd ed., A K Peters, Natick, MA, 1997 Sách, tạp chí
Tiêu đề: Asymptotics and Special Functions
11. F. Stenger, Numerical Methods Based on Sinc and Analytic Functions, Computational Mathematics, vol Sách, tạp chí
Tiêu đề: Numerical Methods Based on Sinc and Analytic Functions
Tác giả: F. Stenger
Nhà XB: Computational Mathematics
12. K. R. Stromberg, An Introduction to Classical Real Analysis, Wadsworth, Belmont, CA, 1981 Sách, tạp chí
Tiêu đề: An Introduction to Classical Real Analysis

TỪ KHÓA LIÊN QUAN