arXiv:2102.08835v2 [cs.GT] 15 May 2021 fdlgtr.Teeoe oedlgto ssf nalrescale. large a on safe converg is probabilities delegation three vote all Therefore, majority, delegators. the of of fraction v fixed of number a total with the conver when delegation Finally, delegato capped non-monotonically. of and nu voting—albeit number free the under the wins if when minority However, degree—occurs ex-ante the capped. lesser alternative is a is, have to can that same—but voters minorities, The favors ex-ante. delegation winning free that show We 02Zrc, Zurich, 8092 E-T n CEPR and CER-ETH eeaievt eeainwe eeaosd o nwteprefer the know not do delegators when delegation vote examine We Z¨urichbergstrasse 18 [email protected] asGersbach Hans oeDlgto ihUcranNme fVoters of Number Uncertain with Delegation Vote atudtd a 8 2021 18, May updated: Last 02Zrc,Switzerland Zurich, 8092 [email protected] kk Mamageishvili Akaki Z¨urichbergstrasse 18 CER-ETH Abstract e oteoeudrconventional under one the to ges sicess h rbblt that probability the increases, rs ooe omte h number the matter no one, to e hthv oe hneof chance lower a have that s [email protected] 02Zrc,Switzerland Zurich, 8092 tr scnegn oinfinity to converging is oters bro oigrgt actual rights voting of mber Z¨urichbergstrasse 18 avrSchneider Manvir ne frepresentatives. of ences CER-ETH 1 Introduction

While a particular type of vote delegation, in the form of in shareholder meetings, has a long tradition, whether and how vote delegation should be allowed in other environments has become a prominent theme in scientific and public discussions. “Liquid ”, for instance, is a concept entailing the right for citizens to delegate voting rights to other citizens. This has been advocated by several political parties and is used internally by the German Pirate party or the Demoex party in . It is implemented in several online platforms such as LiquidFeedback1 and Google Votes2 A further reference is Escoffier B. (2019). It is an important way to govern blockchains, see Goodman (2014) and Damg˚ard et al. (2020). A priori it is not obvious how liquid voting compares to standard voting procedures such as voluntary voting. It is thus useful to compare to standard voting in specific environments. In this paper, we examine how voting outcomes are affected when vote delegation is allowed, but delegators do not know the preferences of the proxy to whom they delegate. In particular, we study the following problem. A polity—say a jurisdiction or a blockchain community—decides between two alternatives A and B. A fraction of individuals does not want to participate in the voting. Henceforth, these individuals are called “delegators”. They abstain in traditional voting and delegate their vote in liquid democracy. There are many reasons why individuals do not want to participate in voting. Either they want to avoid facing the costs of informing themselves or do not want to incur the cost of voting. In the context of blockchains and proof-of-stake, individuals may have the option to delegate their stake to other individuals, thereby participating in the rewards from validating transactions without having to run the validation software themselves. Our central assumption is that delegators do not know the preferences of the individuals who will vote, that is, whether these individuals favor A or B. We call individuals willing to vote “voters” and individuals favoring A (B) “A-voters” (“B-voters”) and the collection of “A-voters” (“B-voters”) party A (party B). From the perspective of delegators, voters themselves and any outsider who might want to design vote delegation, the probability that a voter favors A is some number p (0

1www.liquidfeedback.org (retrieved on February 18, 2021). 2https://www.tdcommons.org/cgi/viewcontent.cgi?article=1092&context=dpubs_series (retrieved on February 18, 2021).

2 to the probability under standard voting. However, the convergence is not monotonic. The results show that outcomes with delegation may significantly differ from standard election outcomes and probabilistically violate the majority principle, i.e. a majority of citizens who is willing to cast a vote for a particular alternative may lose. This result might caution blockchain companies: to implement vote delegation as it might raise the risk that dishonest agents become a majority through vote delegation.

2 Related Literature

A couple of recent papers sheds new light on how vote delegation impacts outcomes. Important studies studying vote delegation from an algorithmic and AI perspective have developed several delegation rules that allow examining how many votes a representative can and should obtain3 and whether delegation votes to neighbours in the network may deliver more desirable outcomes than direct voting4. The result of Kahng et al. (2018) is similar to our result in spirit. The authors show that even with a delegation from a less informed to better informed voter, the probability of implementing the right alternative in a generic network is decreasing. The setting studied in the paper is different, though, as it focuses on the information acquisition aspect of voting, with voters having the same preferences. Delegation games on networks were studied by Bloembergen et al. (2019) and Escoffier B. (2019). Bloembergen et al. (2019) identify conditions for the existence of Nash equilibria of such games, while Escoffier B. (2019) shows that in more general setups, no Nash equilibrium may exist, and it may even be NP -complete to decide whether one exists at all. We adopt a collective decision- framework and study how free delegation or capped delegation impacts outcomes when delegators do not know the preferences of individuals to whom they delegate. models have been studied recently by Pivato and Soh (2020) and Abramowitz and Mattei (2019). Pivato and Soh (2020) study a model where voters can choose any legislator as representative. Hence, a legislator can represent any number of voters, which will be his/her weight. In the actual voting all legislators vote and votes are counted according the weights. In their model they show that for large elections the voting outcome of the legislator’s votes is the same as if all voters voted directly. Soh (2020) uses this model and studies other forms of voting, such as Weighted Approval Voting and Majority Judgement. He shows similar as for Weighted Voting that for large elections the voting outcome is the same as under direct voting. Unlike the literature on liquid democracy, Abramowitz and Mattei (2019) introduce Flexible Rep- resentative Democracy a new model which studies a mixture of representative and where first a set of experts/representatives is elected by the voters and stays fixed during each term. When voting on an issue, a voter faces the decision of how to allocate his/her vote among a subset of the representatives. In addition it may also be possible for voters to vote directly. Comparing this model to our model, we note that random uniform delegation can be achieved by distributing the voting power of a voter uniformly among all representatives. Nevertheless, our model is different from this setting since delegators delegate their votes to any voter in the society. The representatives are the partisans which are not elected.

3 Model

We consider a large polity (a society or a blockchain community) that faces a binary choice between two alternatives A and B. There is a group of m ∈ N+ individuals who do not want to vote and either abstain under conventional voting or delegate their voting rights under vote delegation. The remaining population votes and are thus called “voters”. They have private information about their preference for A or B. Hence, when individuals delegate their voting rights to a voter, this is equivalent to uniformly and randomly delegating a voting right to one voter. From the perspective of other individuals, a

3See G¨olz et al. (2018) and Kotsialou and Riley (2020). 4See Kahng et al. (2018).

3 voter prefers A (B) with probability p (1−p), where 0 2 . Voters favoring A (B) are called “A-voters” (“B-voters”) and the respective group is called “A-party” (“B-party”). We assume that the number of voters is given by a random variable with distribution F , that is, the probability that there are i voters is equal to F (i). We compare three voting processes:

• Conventional voting: Each voter casts one vote.

• Free delegation: m delegators randomly delegate their voting rights. Each voter casts all votes that is corresponding to his/her voting rights.

• Capped delegation: m delegators delegate their voting rights randomly to voters. If a voter has reached the cap, all further voting rights are distributed among the remaining voters up to the cap. If all voters have reached their cap, superfluous voting rights are eliminated.

Q(n,p) denotes the probability that n voters are X-voters, with X ∈ {A, B} and each voter being X-voter with probability p. Then,

∞ i Q(n,p)= F (i) pn(1 − p)i−n. (1) n i n X=   We start with conventional voting and denote by P (p) the probability that A wins. It is calculated by the following formula:

∞ ∞ P (p)= Q(k,p)Q(l, 1 − p)g(k, l), (2) k l X=0 X=0 where g(k, l) is the probability that A-voters are in majority if there are k A-voters and lB-voters, it is calculated in the following way:

1, if k >l, 1 g(k, l) :=  2 , if k = l, 0, if k

 In case of free vote delegation, the probability that A wins, denoted by P (p,m), is equal to:

∞ ∞ P (p,m)= Q(k,p)Q(l, 1 − p)G(k,l,m), (3) Xk=0 Xl=0 where G(k,l,m) denotes the probability that A-voters win if there are k A-voters, l B-voters and m voters delegate their votes. The value is calculated in the following way:

m m k h l m−h G(k,l,m)= g(k + h, l + m − h). h k + l k + l h X=0      We now consider the delegation procedure when delegated votes are distributed randomly among the other voters, with the restriction that no voter obtains more than c votes. If a voter already has c votes, s/he is not allowed to receive more, and, therefore, leaves the pool of receivers. We invalidate the rest of delegated votes, if there are any. With cap c, the probability that A-voters have a majority, denoted as Pc(p,m), is equal to

∞ ∞ Pc(p,m)= Q(k,p)Q(l, 1 − p)Gc(k,l,m), (4) Xk=0 Xl=0

4 where Gc(k,l,m) denotes the probability that A-voters win if there are k A-voters, lB-voters and m voters who delegate their votes, with an individual voter being allowed to have c votes at most. The value is calculated in the following way:

m m k h l m−h Gc(k,l,m)= g(min{k + h, ck}, min{l + m − h, cl}). h k + l k + l Xh=0      Since ck is the maximum amount of votes majority voters can have, we take min{k + h, ck} to calculate the total number of votes for majority voters. Similarly, cl is the maximum amount of votes minority voters can have. Therefore, we take a value min{l + m − h, cl} to calculate the total number of votes minority voters can have. For notational convenience, we denote a binomial random variable with parameters n and p by Bin(n,p), for any n ∈ N and p ∈ [0, 1].

4 Results

To gain an intuition in the formal results, we start with numerical examples. In numerical examples ni we model the number of voters as a Poisson random variable, with average n. That is, F (i) = eni! . To clearly indicate the distribution with parameter n, we occasionally include n in the function P (·). This assumption is a standard tool to make the analysis of voting outcomes analytically tractable. Namely, by decomposition property, the number of A-voters is Poisson random variable with average n · p and the number of B-voters is Poisson random variable with average n · (1 − p).5 Moreover, these two random variables are independent. Poisson games was introduced in Myerson (1998).

n p m P (n,p) P (n,p,m) P2(n,p,m) 20 0.6 1 0.81413 0.808443 0.808443 20 0.6 2 0.81413 0.804256 0.804256 20 0.6 5 0.81413 0.796578 0.796616 20 0.6 10 0.81413 0.791246 0.792627 20 0.6 300 0.81413 0.808516 0.81413

Table 1: Probabilities of A winning under conventional voting P (n,p), free delegation P (n,p,m) and capped delegation Pc(n,p,m) for cap c = 2.

Table 1 reveals that the likelihood of A winning is smaller with free delegation than under con- ventional voting. The same occurs with capped delegation but is slightly less pronounced than with free delegation. We also observe that when m increases, both the winning probabilities for free del- egation and for capped delegation first decline and then start to converge to the probability under conventional voting. The pattern in Table 1 repeats for different values of n and p. In the following, we prove formal results.

4.1 Free Delegation We start with free delegation and show the following:

Theorem 1 P (p,m) < P (p) for any m> 0 and p> 0.5.

5Indeed, this follows directly from

∞ i k ∞ i−k k ∞ i−k k n i − (np) (n(1 − p)) k! i (np) (n(1 − p)) (np) Q(k,p)= pk(1 − p)i k = = = . eni! k enpk! en(1−p)i! k enpk! en(1−p)(i − k)! enpk! i k ! i=0 ! i=0 X= X X

5 In other words, Theorem 1 says that free delegation probabilistically favors the ex-ante minority, since the probability that the minority wins under vote delegation is 1−P (p,m), while the probability that the minority wins in conventional voting is equal to 1 − P (p). Before proving the theorem, we show the following lemma:

Lemma 1 For any l,t,m ∈ N we have G(l + t,l,m)+ G(l, l + t,m) = 1.

Proof of Lemma 1. Similarly to the definition of G, let H(k,l,m) denote the probability that B wins if A voters have k votes, B voters have l votes and m voters are delegating. Then,

G(l + t,l,m)= H(l, l + t,m) = 1 − G(l, l + t,m), (5)

where the first equality follows from the symmetry, and the second equality follows from the fact that events of A and B winning are complementary. With this Lemma we proceed to prove Theorem 1. Proof of Theorem 1. To prove that P (p,m) is smaller than P (p) for m> 0 and p> 0.5, we take a closer look at the summands in P (p,m) and P (p), in (4) and (2), respectively. We consider different cases of k, l and compare the summands of P (p,m) and P (p). Recall the definition of g(k, l). For fixed k > l, the summand in P (p) is equal to

Q(k,p)Q(l, 1 − p), whereas the corresponding summand in P (p,m) is

Q(k,p)Q(l, 1 − p)G(k,l,m) ≤ Q(k,p)Q(l, 1 − p), since G(k,l,m) ≤ 1. That is, the summand in P (p,m) is smaller (or equal) than the summand in P (p) for k > l. Their difference is

Q(k,p)Q(l, 1 − p)(1 − G(k,l,m)) ≥ 0.

1 For k = l, both g(k, l) and G(k,l,m) are equal to 2 and consequently, the corresponding summands in both P (p,m) and P (p) are equal. The only critical case is for fixed k < l. In this case, the summand in P (p) is zero, whereas the summand in P (p,m) is greater than or equal than zero. That is, the summand in P (p,m) is greater than (or equal to) the summand in P (p) for k < l. The difference is

Q(k,p)Q(l, 1 − p)(0 − G(k,l,m)) = −Q(k,p)Q(l, 1 − p)G(k,l,m).

We compare the differences in the cases where k > l and k < l. If the difference in the case k > l is greater than the difference in absolute value in the case k < l, then overall, P (p,m) < P (p). For given m, we consider the following cases: If k ≥ l + m + 1, then g(k, l) = G(k,l,m) =1. If 1 k ≤ l − (m + 1), then g(k, l)= G(k,l,m)=0. If k = l, then g(k, l)= G(k,l,m)= 2 . The 2m relevant cases are k = l ± 1, k = l ± 2, ..., k = l ± m. Let t ∈ {1, 2, ..., m}. If k = l + t, then g(l + t, l)=1 and if k = l − t, then g(l − t, l) = 0. By Lemma 1, we have

G(l + t,l,m)+ G(l, l + t,m) = 1. (6)

We subtract from the terms in P (p) − P (p,m), where k = l + t, the terms in P (p) − P (p,m),

6 where k = l − t, that is

∞ ∞ ½ Q(l, 1 − p)Q(k,p)(g(k, l) − G(k,l,m))½{k=l+t} − Q(l, 1 − p)Q(k,p)(|g(k, l) − G(k,l,m)|) {k=l−t}, l l X=0 X=0 and show that this difference is > 0. Because g(l +t, l)=1and g(l, l +t) = 0, we obtain the following:

∞ ∞ Q(l, 1 − p)Q(l + t,p)(1 − G(l + t,l,m)) − Q(l, 1 − p)Q(l − t,p)G(l − t,l,m). Xl=0 Xl=0 We simplify this by rearranging terms and using (6). In the first sum, we only use (6). In the second sum we change the variable l → l + t and obtain

∞ ∞ Q(l, 1 − p)Q(l + t,p)G(l, l + t,m) − Q(l + t, 1 − p)Q(l,p)G(l, l + t,m) l=0 l=0 X∞ X = (Q(l, 1 − p)Q(l + t,p) − Q(l + t, 1 − p)Q(l,p)) G(l, l + t,m). l X=0 =:∆ | {z } As G(l, l + t,m) ≥ 0, we need to show that ∆ > 0. From (1), we have the following:

∞ i Q(l + t,p)= F (i) (1 − p)l+tpi−l−t, l + t i=Xl+t   ∞ i Q(l, 1 − p)= F (i) (1 − p)lpi−l, l Xi=l   ∞ i Q(l + t, 1 − p)= F (i) (1 − p)l+tpi−l−t, l + t i=Xl+t   ∞ i Q(l,p)= F (i) pl(1 − p)i−l. l Xi=l   ∞ ∞ The Cauchy product rule states that for two absolute converging sums k=0 ak and k=0 bk, ∞ n their product is equal to ck where ck = akbn k. The product of Q(l + t,p) and Q(l, 1 − p) n=0 k=0 − P P is equal to P P ∞ ∞ i i Q(l + t,p) · Q(l, 1 − p)= F (i) pl+t(1 − p)i−l−t F (i) (1 − p)lpi−l l + t l  i=0    i=0    ∞Xn X k n − k = F (k) pl+t(1 − p)k−l−tF (n − k) (1 − p)lpn−k−l l + t l n=0 k=0     X∞ Xn k n − k = F (k)F (n − k) pn−k+t(1 − p)k−t. l + t l n=0 X Xk=0   

7 Similarly, the product of Q(l + t, 1 − p) and Q(l,p) is equal to

∞ ∞ i i Q(l + t, 1 − p) · Q(l,p)= F (i) (1 − p)l+tpi−l−t F (i) pl(1 − p)i−l l + t l  i=0    i=0    ∞Xn X k n − k = F (k) (1 − p)l+tpk−l−tF (n − k) pl(1 − p)n−k−l l + t l n=0 k=0     X∞ Xn k n − k = F (k)F (n − k) pk−t(1 − p)n−k+t. l + t l n X=0 Xk=0    Hence, the difference ∆ is given by

∞ n k n − k ∆= F (k)F (n − k) pn−k+t(1 − p)k−t − (1 − p)n−k+tpk−t . l + t l n=0 k=0    X X   ∆ > 0 if pn−k+t(1 − p)k−t − (1 − p)n−k+tpk−t > 0. The latter condition is equivalent to p> 1/2. This holds for all relevant cases (t ∈ {1, 2, ..., m}). As the difference to 1 in the cases where k is larger than l in P (p,m) is greater than the difference to 0 in the cases where k is smaller than l, it follows that P (p,m) < P (p) for m> 0,p> 0.5.

4.2 Capped Delegation We establish the following results in the case of capped delegation:

Theorem 2 Pc(p,m) < P (p) for any m> 0, c> 0 and p> 0.5.

Theorem 3 P (p,m) < Pc(p,m) for any m> 0, c> 0 and p> 0.5.

Theorem 2 shows that capped delegation probabilistically favors the ex-ante minority. On the other hand, Theorem 3 shows that capped delegation is better for the ex-ante majority than free delegation. Together, they imply that capped delegation is in-between standard voting (no delegation) and free delegation, with respect to the probability of majority winning. The latter also confirms the observation obtained in the numerical example. Before proving the theorems, we need to prove the following two lemmata:

Lemma 2 For c ≥ 2, l,m ∈ N and t ∈ {1, ..., m}:

Gc(l + t,l,m)+ Gc(l, l + t,m) = 1.

Proof of Lemma 2. Similarly to the definition of Gc, let Hc(k,l,m) denote the probability that B wins if A voters have k votes, B voters have l votes, m voters are delegating and the cap is equal to c. Then,

Gc(l + t,l,m)= Hc(l, l + t,m) = 1 − Gc(l, l + t,m), (7)

where the first equality follows from the symmetry, and the second equality follows from the fact that events of A and B winning are complementary.

1 N Lemma 3 Gc(l,l,m)= 2 for any c,l,m ∈ .

8 Proof of Lemma 3. Proof follows immediately from the symmetry. Next, we prove Theorem 2. Proof of Theorem 2. As for the proof of Theorem 1, the key lies in Lemma 2. We proceed as in 1 the proof of Theorem 1. Recall that g(k, l)=1if k > l, g(k, l) = 2 if k < l and g(k, l)=0if k = l. That is why, for fixed k > l a representative summand in P (p) is equal to

Q(k,p)Q(l, 1 − p), whereas a representative summand of Pc(p,m), in (4), is

Q(k,p)Q(l, 1 − p)Gc(k,l,m) ≤ Q(k,p)Q(l, 1 − p), since Gc(k,l,m) ≤ 1. Therefore, the summand in Pc(p,m) is smaller (or equal) than the summand in P (p) for k > l. The difference is

Q(k,p)Q(l, 1 − p)(1 − Gc(k,l,m)) ≥ 0.

1 For fixed k = l, both g(k, l)= Gc(k,l,m)= 2 by Lemma 3. Consequently, the summands of both Pc(p,m) and P (p) are equal. The only critical case is for fixed k < l. In that case, the summand in P (p) is zero, whereas the summand in Pc(p,m) is greater or equal than zero. So the summand in Pc(p,m) is greater than (or equal to) the summand in P (p) for k < l. The difference of the two summands is equal to

Q(k,p)Q(l, 1 − p)(0 − Gc(k,l,m)) = −Q(k,p)Q(l, 1 − p)Gc(k,l,m).

We have to compare the differences in the cases where k > l and k < l. If the difference in the case where k > l is greater than the difference in absolute value in the case where k < l, then overall, Pc(p,m) < P (p). For given m, we need to take care of the following: If k ≥ l + m + 1 then g(k, l)= Gc(k,l,m) = 1. 1 If k ≤ l − (m + 1), then g(k, l) = Gc(k,l,m)=0. If k = l, then g(k, l) = Gc(k,l,m) = 2 . The 2m relevant cases are k = l ± 1, l ± 2, ..., l ± m. Let t ∈ {1, 2, ..., m}. If k = l + t, then g(l + t, l)=1 and if k = l − t, then g(l − t, l) = 0. By Lemma 2 we have

Gc(l + t,l,m)+ Gc(l, l + t,m) = 1. (8)

We subtract from the terms in P (p) − Pc(p,m), where k = l + t, the terms in P (p) − Pc(p,m), where k = l − t, that is

∞ ∞ ½ Q(l, 1 − p)Q(k,p)(g(k, l) − Gc(k,l,m))½{k=l+t} − Q(l, 1 − p)Q(k,p)(|g(k, l) − Gc(k,l,m)|) {k=l−t}, Xl=0 Xl=0 and show that this difference is > 0. Because g(l +t, l)=1and g(l, l +t) = 0, we obtain the following:

∞ ∞ Q(l, 1 − p)Q(l + t,p)(1 − Gc(l + t,l,m)) − Q(l, 1 − p)Q(l − t,p)Gc(l − t,l,m). Xl=0 Xl=0 We simplify this by rearranging terms and using (8). In the first sum, we only use (8). In the second sum we change the variable l → l + t and obtain

9 ∞ ∞ Q(l, 1 − p)Q(l + t,p)Gc(l, l + t,m) − Q(l + t, 1 − p)Q(l,p)Gc(l, l + t,m) l=0 l=0 X∞ X = (Q(l, 1 − p)Q(l + t,p) − Q(l + t, 1 − p)Q(l,p)) Gc(l, l + t,m). l X=0 =:∆ | {z } As Gc(l, l + t,m) ≥ 0, we need to show that ∆ > 0. ∞ ∞ The Cauchy product rule states that for two absolute converging sums k=0 ak and k=0 bk, ∞ n their product is equal to ck where ck = akbn k. The product of Q(l + t,p) and Q(l, 1 − p) n=0 k=0 − P P is equal to P P ∞ ∞ i i Q(l + t,p) · Q(l, 1 − p)= F (i) pl+t(1 − p)i−l−t F (i) (1 − p)lpi−l l + t l  i=0    i=0    ∞Xn X k n − k = F (k) pl+t(1 − p)k−l−tF (n − k) (1 − p)lpn−k−l l + t l n=0 k=0     X∞ Xn k n − k = F (k)F (n − k) pn−k+t(1 − p)k−t. l + t l n=0 X Xk=0    Similarly, the product of Q(l + t, 1 − p) and Q(l,p) is equal to

∞ ∞ i i Q(l + t, 1 − p) · Q(l,p)= F (i) (1 − p)l+tpi−l−t F (i) pl(1 − p)i−l l + t l  i=0    i=0    ∞Xn X k n − k = F (k) (1 − p)l+tpk−l−tF (n − k) pl(1 − p)n−k−l l + t l n=0 k=0     X∞ Xn k n − k = F (k)F (n − k) pk−t(1 − p)n−k+t. l + t l n=0 k X X=0    Hence, the difference ∆ is given by

∞ n k n − k ∆= F (k)F (n − k) pn−k+t(1 − p)k−t − (1 − p)n−k+tpk−t . l + t l n=0 k=0    X X   ∆ > 0 if pn−k+t(1 − p)k−t − (1 − p)n−k+tpk−t > 0. The latter condition is equivalent to p> 1/2. This holds for all relevant cases (t ∈ {1, 2, ..., m}). As the difference to 1 in the cases where k is larger than l in Pc(p,m) is greater than the difference to 0 in the cases where k is smaller than l, it follows that Pc(p,m) < P (p) for m> 0,p> 0.5.

Proof of Theorem 3. The proof is analogous to the proof of Theorem 2.

4.3 Asymptotic Behavior of Delegation In this section, we consider three cases of asymptotic behavior of delegation. In the first, the prob- ability that a voter is in the majority is large. In the second, the average total number of voters is large. Both cases describe typical assumptions in real-world situations. In the first, one assumes there are very few voters in the minority, for example malicious voters in the context of blockchains. In the second, we look at a large electorate, with fixed shares of majority and minority voters. In the

10 third, we show the convergence of both delegation rules towards conventional voting as the number of delegators goes to infinity. First, we obtain a rather straightforward proposition:

Proposition 1 limp→1 P (p,m) = 1 for any fixed n and m.

Proof of Proposition 1. We can verify that P (1,m) = 1.

∞ m k m 0 0 P (1,m)= Q(k, 1) g(k + m, 0) = 1. m k k Xk=0      Indeed, only the terms where l = 0 and h = m survive. Since the function P (p,m) is uniformly continuous in p, the proposition is proved. Next, we consider a large electorate. It is straightforward that with a constant number of delega- tors, the probability that the majority wins converges to one as the total population size converges to infinity. We show that the same holds even if there are arbitrarily many delegators.

1 Proposition 2 limn→∞ P (n,p,m) = 1 for any fixed p> 2 and any m, where m can even depend on n.

Proof of Proposition 2. 1 Let p> 0.5 and ǫ ∈ (0,p − 2 ). Define two Poisson random variables K, with parameter np, and L, 1 with parameter n(1−p). The probability that A-voters have at least n( 2 +ǫ) votes can be bounded by using the Poisson random variable concentration inequalities from Mitzenmacher and Upfal (2005):

e−λ(eλ)x P (X ≤ x) ≤ (9) xx and

e−λ(eλ)x P (X ≥ x) ≤ . (10) xx 1 Using (9) for n( 2 + ǫ) and np, we obtain: n 1 +ǫ p 2 n( 1 +ǫ) 1 (enp) 2 1+ǫ n→∞ P [K

p q < ep−q, q  

1 p−q q where q := 2 + ǫ. Consider f(p) = e − (p/q) . Then, f(q)= 0 and f is increasing in p. The ∂f(p) p−q q−1 p−q q−1 latter holds because ∂p = e − (p/q) > 0 for any p>q, since e > 1 and (p/q) < 1. 1 Hence, the probability that A-voters have at least n( 2 + ǫ) votes is

1 1 n→∞ P [K ≥ n( + ǫ)] = 1 − P [K

11 1 At the same time we apply (10) for n( 2 − ǫ) and n(1 − p), we obtain:

n( 1 −ǫ) 1 (en(1 − p)) 2 n→∞ P [L>n( − ǫ)] ≤ 1 −→ 0. (11) 2 n(1−p) 1 n( 2 −ǫ) e (n( 2 − ǫ))

The last implication follows from the following: We can rewrite the fraction on the right-hand side as follows

n( 1 −ǫ)(1+log(n(1−p))) e 2 1 1 − p 1 − = exp n( − ǫ) 1 + log(n(1 − p)) − − log(n( − ǫ)) . n( 1 −ǫ)( (1 p) +log(n( 1 −ǫ))) 2 1 2 2 ( 1 −ǫ) 2 2 − ǫ e 2    =:g | {z } Note that g is independent of n as we can write

1 − p 1 − p g = 1 − 1 + log( 1 ). 2 − ǫ 2 − ǫ

By assumption on p and ǫ we have that 1 − p 1 < 1. 2 − ǫ We consider the following function for any y ∈ (0, 1),

f(y) := 1 − y + log(y).

1−p Then f( 1 )= g. Function f has the following properties: First, limy→0 f(y)= −∞ and limy→1 f(y)= 2 −ǫ ′ 1 0. Second, the derivative f (y)= −1+ y > 0 since y < 1. Hence, f(y) is negative for any y < 1. This 1 implies that g < 0 and hence the right-hand side of Equation (11) is exp(n( 2 − ǫ)g) and converges to 0 for n →∞ since g is negative. 1 Hence, the probability that B-voters have at most n( 2 − ǫ) votes is

1 1 n→∞ P [L ≤ n( − ǫ)] = 1 − P [L>n( − ǫ)] −→ 1. 2 2

These two results and the fact that K,L are independent yield:

1 1 1 1 n P [K ≥ n( + ǫ) and L ≤ n( − ǫ)] = P [K ≥ n( + ǫ)] P [L ≤ n( − ǫ)] −→→∞ 1. 2 2 2 2 n→∞ n→∞ −→ 1 −→ 1 | {z } | {z } It follows that for m< 2ǫn, A-voters are in the majority with probability 1, if n goes to infinity. Next, we consider m > 2ǫn. From above, we know that for high enough n, with probability 1 1 1 K ≥ n( 2 + ǫ) and L ≤ n( 2 − ǫ). Hence,

1 K 2 + ǫ ≥ 1 > 1. L 2 − ǫ

We define the i.i.d. random variables:

12 1 +1 w.p. 2 + ǫ Xi := 1 (−1 w.p. 2 − ǫ for i = 1, · · · ,m. The surplus of delegated votes to party A-voters is the sum of the Xi. We know 1 m m→∞ that E[Xi] = 2ǫ> 0 and by the law of large numbers, m i=1 Xi −→ 2ǫ> 0. As m> 2ǫn, n →∞ implies m →∞. P We show the same result with capped delegation.

1 N Proposition 3 limn→∞ Pc(n,p,m) = 1 for any fixed p> 2 , any c ∈ and any m, where m can even depend on n.

Proof of Proposition 3. The proof is a direct corollary of Proposition 2 and Theorem 3. Next, we show that the probability that A wins with free delegation converges to the probability that A wins with conventional voting, if m goes to infinity.

Proposition 4 limm→∞ P (p,m)= P (p) for any p ∈ (0, 1).

6 Proof of Proposition 4. First, using Hoeffding’s inequality , we can show that limm→∞ G(k,l,m)= g(k, l) for any fixed k and l. • k 1 Consider fixed pair k, l, so that k = l. Then, with q := k+l = 2 , we have

m m h m−h 1 m m m G(k,l,m)= q (1 − q) + q 2 (1 − q) 2 h 2 m h=⌊ m +1⌋    2  X2 m 1 m m 1 m+1 m = + . 2 h 2 m   h=⌊ m +1⌋      2  X2

Note that the latter summand is zero if m is odd.

– If m is odd:

m 1 m m 1 m 2m 1 G(k,l,m)= = = . 2 h 2 2 2   h=⌊ m +1⌋     X2

– If m is even:

m 1 m m 1 m 1 m 2m 1 G(k,l,m)= + = = . 2  h 2 m  2 2 2   h=⌊ m +1⌋     2    X2   1 Hence, for k = l, we have G(k,l,m)= g(k, l)= 2 . • k 1 Consider fixed pair k, l, so that k < l. Then, with q := k+l < 2 , we have

m m 1 m l−k+m m l−k+m h m−h 2 − 2 G(k,l,m)= q (1 − q) + l−k+m q (1 − q) . l−k m h 2 2 h=⌊ + +1⌋     X2 6See Hoeffding (1963).

13 l−k+m Define a := 2 . The first summand above is equal to the probability P [X > a] for a random variable X ∼ Bin(m,q). We consider P [X ≥ a], which is G(k,l,m) plus a non-negative term (the second term of G(k,l,m)) and show that this converges to 0 and hence also G(k,l,m) converges to 0. Hoeffding’s inequality yields:

P [X ≥ m(q + ǫ)] ≤ exp(−2ǫ2m).

We calculate ǫ by solving m(q + ǫ)= a:

l − k + m ǫ = − q. 2m

Let t := l − k, then by Hoeffding’s inequality:

l − k + m P [X ≥ a] ≤ exp(−2( − q)2m) 2m 1 t2 = exp(−m(2q2 − 2q + ) − t − + 2qt). 2 2m

2 1 The right-hand side goes to 0 for m → ∞ if 2q − 2q + 2 > 0. This inequality is true for any 1 1 q 6= 2 . As q < 2 , the RHS goes to 0 for m →∞. That is,

lim P [X ≥ a] = 0. m→∞

Hence, for k < l we have limm→∞ G(k,l,m)= g(k, l) = 0. • k 1 Consider fixed pair k, l, so that k > l. Then, with q := k+l > 2 we have

m m 1 m l−k+m m l−k+m h m−h 2 − 2 G(k,l,m)= q (1 − q) + l−k+m q (1 − q) . l−k m h 2 2 h=⌊ + +1⌋     X2

l−k+m Define a := 2 . The first summand above is equal to the probability P [X > a] for a random variable X ∼ Bin(m,q). As before, we consider the value P [X ≥ a]. Hoeffding’s inequality yields:

P [X

We calculate ǫ by solving m(q − ǫ)= a:

l − k + m ǫ = q − . 2m

Let t := |l − k|, then, by Hoeffding’s inequality:

l − k + m P [X < a] ≤ exp(−2(q − )2m) 2m 1 t2 = exp(−m(2q2 − 2q + ) − t − + 2qt). 2 2m

14 2 1 The RHS of the latter converges to 0 for m →∞ if 2q − 2q + 2 > 0. This inequality is true for 1 1 any q 6= 2 . As q > 2 , the RHS converges to 0 for m →∞. That is,

lim P [X < a] = 0. m→∞

Therefore, lim P [X ≥ a] = lim (1 − P [X < a]) = 1. m→∞ m→∞

Hence, for k > l, we have limm→∞ G(k,l,m)= g(k, l) = 1.

It follows that limm→∞ G(k,l,m)= g(k, l) for any fixed k and l.

In the next step, we want to show that limm→∞ P (p,m)= P (p). Define,

ak(m) := Q(k,p)Q(l, 1 − p)G(k,l,m).

Then, ∞ P (p,m)= ak(m). k X=0 Note that, as G(k,l,m) ≤ 1,

∞ |ak(m)|≤ Q(k,p) Q(l, 1 − p)= Q(k,p) =: Mk, Xl=0 ∞ and k=0 Mk = 1 < ∞. Then, by dominated convergence theorem,

P ∞ lim P (p,m)= lim ak(m) m→∞ m→∞ k=0 X∞ ∞ = Q(k,p) lim Q(l, 1 − p)G(k,l,m). m→∞ Xk=0 Xl=0 Define bl(m) := Q(l, 1 − p)G(k,l,m).

Note that |bl(m)|≤ Q(l, 1 − p) =: Ml

∞ and l=0 Ml = 1 < ∞. Hence, by dominated convergence theorem,

P ∞ ∞ lim bl(m)= lim bl(m) m→∞ m→∞ l=0 l=0 X X∞ = Q(l, 1 − p) lim G(k,l,m). m→∞ Xl=0

15 Taken together, by applying dominated convergence theorem twice, we obtain:

∞ lim P (p,m)= lim ak(m) m→∞ m→∞ k=0 X∞ ∞ = Q(k,p) lim bl(m) m→∞ k=0 l=0 X∞ ∞ X = Q(k,p) lim bl(m) m→∞ k=0 l=0 X∞ X∞ = Q(k,p) Q(l, 1 − p) lim G(k,l,m) m→∞ k=0 l=0 X∞ X∞ = Q(k,p) Q(l, 1 − p)g(k, l)= P (p). Xk=0 Xl=0

Similarly, for capped delegation, we show:

Proposition 5 limm→∞ Pc(p,m)= P (p) for any c ∈ N and p ∈ (0, 1). Proof of Proposition 5. The proof is an immediate corollary of Proposition 4 and Theorem 3.

5 Discussion and Conclusion

We showed that the introduction of vote delegation leads to lower probabilities of winning for the ex-ante majority. Capped delegation leads to even lower probabilities than free delegation. However, both delegation processes lead to lower probabilities than conventional voting. These results are particularly important in blockchain governance if one does not want dishonest agents to increase their probability of winning. Although the setting we analyze in the paper is, as simple as possible, we already obtain non-trivial observations. In addition to the formal results, we also conjecture that P (n,p,m) is first decreasing and then increasing in m, based on numerical evidence. This conjecture is interesting in the context of delegation as a strategic decision. For the delegators, whether or not they belong to the ex-ante majority, choosing between delegating and abstaining becomes a strategic decision. If we prove only one part of the conjecture, namely that the winning probability is increasing from some point, we will get that in the equilibrium all majority voters are delegating for a large enough number of delegators m. First, note that nobody delegating can not be an equilibrium state, because minority voters prefer to delegate. Once some of them delegate, the probability that the majority of voters winning is smaller than the one in conventional voting. Using Proposition 2, we get that with a large number of delegators, all majority voters delegating is sustainable in the equilibrium. We leave this challenging issue to future research.

Acknowledgements This research was partially supported by the Zurich Information Security and Privacy Center (ZISC).

References

Abramowitz, B. and Mattei, N. (2019). Flexible Representative Democracy: An Introduction with Binary Issues. In Proceedings of the Twenty-Eighth International Joint Conference on Artificial

16 Intelligence, IJCAI-19, pages 3–10. International Joint Conferences on Artificial Intelligence Orga- nization.

Bloembergen, D., Grossi, D., and Lackner, M. (2019). On Rational Delegations in Liquid Democracy. Proceedings of the AAAI Conference on Artificial Intelligence, 33(1):1796–1803.

Damg˚ard, I., Gersbach, H., Maurer, U., Nielsen, J. B., Orlandi, C., and Pedersen, T. P. (2020). Concordium White Paper.

Escoffier B., Gilbert H., P.-L. A. (2019). The Convergence of Iterative Delegations in Liquid Democ- racy in a Social Network. In: Fotakis D., Markakis E. (eds) Algorithmic Game Theory. SAGT 2019. Lecture Notes in Computer Science, 11801:284–297.

G¨olz, P., Kahng, A., Mackenzie, S., and Procaccia, A. D. (2018). The Fluid Mechanics of Liquid Democracy. In: Christodoulou G., and Harks T. (eds) Web and Internet Economics - 14th Inter- national Conference, WINE 2018, Proceedings. Lecture Notes in Computer Science, 11316:188–202.

Goodman, L. (2014). Tezos – Self-amending Crypto-ledger. White Paper.

Hoeffding, W. (1963). Probability Inequalities for Sums of Bounded Random Variables. Journal of the American Statistical Association, 58(301):13–30.

Kahng, A., Mackenzie, S., and Procaccia, A. (2018). Liquid Democracy: An Algorithmic Perspective. Proceedings of the AAAI Conference on Artificial Intelligence, 32(1):1095–1102.

Kotsialou, G. and Riley, L. (2020). Incentivising Participation in Liquid Democracy with Breadth- First Delegation. Proceedings of the 19th International Conference on Autonomous Agents and MultiAgent Systems, page 638–644.

Mitzenmacher, M. and Upfal, E. (2005). Probability and Computing: Randomized Algorithms and Probabilistic Analysis. Cambridge University Press, page 97.

Myerson, R. (1998). Population Uncertainty and Poisson Games. International Journal of Game Theory, 27:375–392.

Pivato, M. and Soh, A. (2020). Weighted Representative Democracy. Journal of Mathematical Economics, 88:52–63.

Soh, A. (2020). Approval Voting & Majority Judgment in Weighted Representative Democracy. THEMA Working Paper n2020-15 CY Cergy Paris Universite, France.

17