+ All Categories
Home > Documents > First carrot, then stick: how the adaptive hybridization of...

First carrot, then stick: how the adaptive hybridization of...

Date post: 06-Aug-2020
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
36
rsif.royalsocietypublishing.org Research Cite this article: Chen X, Sasaki T, Bra ¨nnstro ¨m A ˚ , Dieckmann U. 2015 First carrot, then stick: how the adaptive hybridization of incentives promotes cooperation. J. R. Soc. Interface 12: 20140935. http://dx.doi.org/10.1098/rsif.2014.0935 Received: 20 August 2014 Accepted: 3 November 2014 Subject Areas: biomathematics, computational biology Keywords: punishment, reward, public good, evolutionary game, social design Author for correspondence: Tatsuya Sasaki e-mail: [email protected] These authors contributed equally to this study. Electronic supplementary material is available at http://dx.doi.org/10.1098/rsif.2014.0935 or via http://rsif.royalsocietypublishing.org. First carrot, then stick: how the adaptive hybridization of incentives promotes cooperation Xiaojie Chen 1,2,† , Tatsuya Sasaki 1,3,† ,A ˚ ke Bra ¨nnstro ¨m 1,4 and Ulf Dieckmann 1 1 Evolution and Ecology Program, International Institute for Applied Systems Analysis (IIASA), Laxenburg 2361, Austria 2 School of Mathematical Sciences, University of Electronic Science and Technology of China, Chengdu 611731, People’s Republic of China 3 Faculty of Mathematics, University of Vienna, Vienna 1090, Austria 4 Department of Mathematics and Mathematical Statistics, Umea ˚ University, Umea ˚ 90187, Sweden TS, 0000-0002-4635-1389; UD, 0000-0001-7089-0393 Social institutions often use rewards and penalties to promote cooperation. Providing incentives tends to be costly, so it is important to find effective and efficient policies for the combined use of rewards and penalties. Most studies of cooperation, however, have addressed rewarding and punishing in isolation and have focused on peer-to-peer sanctioning as opposed to institutional sanctioning. Here, we demonstrate that an institutional sanc- tioning policy we call ‘first carrot, then stick’ is unexpectedly successful in promoting cooperation. The policy switches the incentive from rewarding to punishing when the frequency of cooperators exceeds a threshold. We find that this policy establishes and recovers full cooperation at lower cost and under a wider range of conditions than either rewards or penalties alone, in both well-mixed and spatial populations. In particular, the spatial dynamics of cooperation make it evident how punishment acts as a ‘booster stage’ that capitalizes on and amplifies the pro-social effects of rewarding. Together, our results show that the adaptive hybridization of incentives offers the ‘best of both worlds’ by combining the effectiveness of rewarding in establishing cooperation with the effectiveness of punishing in recovering it, thereby providing a surprisingly inexpensive and widely applicable method of promoting cooperation. 1. Introduction Cooperation is desirable whenever groups of cooperating individuals can reap higher benefits than groups of individuals acting out of individual self-interest. Promoting cooperation can be difficult, however, because a single non-cooperating individual (‘defector’) in a group of cooperators often achieves a higher net benefit by free-riding on the others’ contributions. An efficient policy for promoting cooperation needs to overcome two fundamental challenges: to ensure that cooperators can gain a foothold in a population of defectors and to protect a population of cooperators from exploitation by defectors once cooperation has been established. Incentives can help overcome these challenges [1–3]. The promise of reward or the threat of punishment can induce cooperation among self-interested indi- viduals who would otherwise prefer actions that undermine the public good. At first glance, there might seem to be little difference between a reward and a penalty: after all, cooperation is induced whenever the size of the incentive exceeds the pay-off difference between a cooperator and a defector, regardless of whether the incentive is positive or negative [4]. This equivalence ceases to hold, however, when one considers the costs of implementing an incentive & 2014 The Authors. Published by the Royal Society under the terms of the Creative Commons Attribution License http://creativecommons.org/licenses/by/4.0/, which permits unrestricted use, provided the original author and source are credited.
Transcript
Page 1: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

rsif.royalsocietypublishing.org

ResearchCite this article: Chen X, Sasaki T, Brannstrom

A, Dieckmann U. 2015 First carrot, then stick:

how the adaptive hybridization of incentives

promotes cooperation. J. R. Soc. Interface 12:

20140935.

http://dx.doi.org/10.1098/rsif.2014.0935

Received: 20 August 2014

Accepted: 3 November 2014

Subject Areas:biomathematics, computational biology

Keywords:punishment, reward, public good, evolutionary

game, social design

Author for correspondence:Tatsuya Sasaki

e-mail: [email protected]

†These authors contributed equally to this

study.

Electronic supplementary material is available

at http://dx.doi.org/10.1098/rsif.2014.0935 or

via http://rsif.royalsocietypublishing.org.

& 2014 The Authors. Published by the Royal Society under the terms of the Creative Commons AttributionLicense http://creativecommons.org/licenses/by/4.0/, which permits unrestricted use, provided the originalauthor and source are credited.

First carrot, then stick: how the adaptivehybridization of incentives promotescooperation

Xiaojie Chen1,2,†, Tatsuya Sasaki1,3,†, Ake Brannstrom1,4 and Ulf Dieckmann1

1Evolution and Ecology Program, International Institute for Applied Systems Analysis (IIASA), Laxenburg2361, Austria2School of Mathematical Sciences, University of Electronic Science and Technology of China, Chengdu 611731,People’s Republic of China3Faculty of Mathematics, University of Vienna, Vienna 1090, Austria4Department of Mathematics and Mathematical Statistics, Umea University, Umea 90187, Sweden

TS, 0000-0002-4635-1389; UD, 0000-0001-7089-0393

Social institutions often use rewards and penalties to promote cooperation.

Providing incentives tends to be costly, so it is important to find effective

and efficient policies for the combined use of rewards and penalties. Most

studies of cooperation, however, have addressed rewarding and punishing

in isolation and have focused on peer-to-peer sanctioning as opposed to

institutional sanctioning. Here, we demonstrate that an institutional sanc-

tioning policy we call ‘first carrot, then stick’ is unexpectedly successful in

promoting cooperation. The policy switches the incentive from rewarding

to punishing when the frequency of cooperators exceeds a threshold. We

find that this policy establishes and recovers full cooperation at lower cost

and under a wider range of conditions than either rewards or penalties

alone, in both well-mixed and spatial populations. In particular, the spatial

dynamics of cooperation make it evident how punishment acts as a ‘booster

stage’ that capitalizes on and amplifies the pro-social effects of rewarding.

Together, our results show that the adaptive hybridization of incentives

offers the ‘best of both worlds’ by combining the effectiveness of rewarding

in establishing cooperation with the effectiveness of punishing in recovering

it, thereby providing a surprisingly inexpensive and widely applicable

method of promoting cooperation.

1. IntroductionCooperation is desirable whenever groups of cooperating individuals can reap

higher benefits than groups of individuals acting out of individual self-interest.

Promoting cooperation can be difficult, however, because a single non-cooperating

individual (‘defector’) in a group of cooperators often achieves a higher net benefit

by free-riding on the others’ contributions. An efficient policy for promoting

cooperation needs to overcome two fundamental challenges: to ensure that

cooperators can gain a foothold in a population of defectors and to protect a

population of cooperators from exploitation by defectors once cooperation has

been established.

Incentives can help overcome these challenges [1–3]. The promise of reward

or the threat of punishment can induce cooperation among self-interested indi-

viduals who would otherwise prefer actions that undermine the public good.

At first glance, there might seem to be little difference between a reward and

a penalty: after all, cooperation is induced whenever the size of the incentive

exceeds the pay-off difference between a cooperator and a defector, regardless

of whether the incentive is positive or negative [4]. This equivalence ceases to

hold, however, when one considers the costs of implementing an incentive

Page 2: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

rsif.royalsocietypublishing.orgJ.R.Soc.Interface

12:20140935

2

scheme. Rewarding a large number of cooperators or penaliz-

ing a large number of defectors is either very costly or

becomes ineffective when a limited budget for incentives is

stretched out too far. Pamela Oliver exemplifies this with

the problem of fundraising: ‘If only 5% of the population

needs to contribute to an Arts Fund for it to be successful,

they can be rewarded by having their names printed in a pro-

gram: it would be silly and wasteful to try to punish the 95%

who did not contribute’ [5, p. 125]. While the challenges of

implementing positive and negative incentives are separately

well known [2,3] and weight has traditionally been given to

peer-to-peer punishment [6–8], no study to date has estab-

lished how such incentives should best be combined at an

institutional level to promote cooperation.

Here, we demonstrate how an institution implementing

incentives can effectively establish and recover cooperation at

low cost. Institutional sanctioning is widespread [1,4,9–22];

however, surprisingly few theoretical studies have thus far

considered the effects of institutionalized incentives on the

evolution of cooperation, and the few studies that exist have

either considered rewarding and punishing in isolation

[4,9,19] or did not consider how optional incentives change

with the frequency of cooperators [10–12]. Indeed, sanctioning

agents, such as officers and managers, often alter rewards and

penalties as events unfold. We address this question in an

established game-theoretical framework for studying the evol-

ution of cooperation under institutionalized incentives [4,19].

By considering the strengths of positive and negative incen-

tives as independent variables, we can encompass a range of

hybrid incentive policies. In particular, by allowing the relative

allocation of incentives to rewarding and punishing to vary

with the frequency of cooperators, our framework includes

hybrid incentive policies controlled by adaptive feedback

from the population’s current state.

2. Model and methods2.1. Institutional incentivesWe aim to determine the best way to allocate a budget available

to an institution for promoting cooperation through positive

and negative incentives. As criteria for assessing the per-

formance of alternative sanctioning policies, we consider their

effectiveness and efficiency in promoting cooperation. For

measuring effectiveness, we assess the diversity of conditions

for which full cooperation can be established or recovered

with certainty, and for measuring efficiency, we determine the

cumulative cost and total time required to convert a population

of defectors to full cooperation or to recover full cooperation

after the invasion of a single defector.

2.2. Public good games with dynamic incentivesOur model is based on the public good game for cooperation (C)

and defection (D), widely recognized as the most suitable math-

ematical metaphor for studying cooperation in large groups

[23–28]. We posit well-mixed populations of interacting indi-

viduals. From time to time, individuals randomly selected

from the population form an n-player group with n � 2. A coop-

erator invests a fixed amount c . 0 into a common pool,

whereas a defector invests nothing. The total contribution to

the pool is then multiplied by a public-benefit factor r . 1 and

distributed equally among all n group members. The infamous

‘tragedy of the commons’ [29] arises when r , n and no incen-

tives are applied, because single individuals can then improve

their pay-offs by withholding their contributions.

The total budget for providing incentives is given by ndper group, where d . 0 is the average per capita incentive.

This budget nd is then divided into two parts based on a rela-

tive weight w with 0 � w � 1. The part wnd is equally shared

among the nC cooperators in the group (see Chen et al. [17] for

a similar application to the n-person volunteer’s dilemma),

who thus each obtain a reward awnd/nC, while the remainder

(1 2 w)nd is used for equally punishing the n 2 nC defectors,

who thus each have their pay-offs reduced by b(1 2 w)nd/

(n 2 nC). The factors a,b . 0 are the respective leverages of

rewarding and punishing, i.e. the factors by which a recipient’s

pay-off is increased or decreased relative to the cost of

implementing the incentive. ‘Antisocial’ incentives, rewarding

defectors or punishing cooperators [30], could in principle be

considered, but as such incentives only reduce cooperation

and promote defection, they are not studied here.

We account for feedback from the population’s state by

allowing the weight w to depend on the frequency of coop-

erators. Pure rewarding and pure punishing correspond to

w ¼ 1 or w ¼ 0, respectively. Therefore, a cooperator and a

defector obtain the pay-offs

rcnC

n� cþ awnd

nCand

rcnC

n� b(1� w)nd

n� nC, (2:1)

respectively.

2.3. Replicator dynamicsWe assume replicator dynamics [31], which describe how the

frequencies of different strategies change in infinitely large,

well-mixed populations. Replicator dynamics are governed

by a system of differential equations, _xi ¼ xi(Pi � �P), in

which xi, Pi and �P denote, respectively, the frequency of strat-

egy i, the average pay-off for strategy i, and the average pay-off

in the whole population ( �P ¼P

ixiPi, withP

ixi ¼ 1). In the

public good game studied here, we consider cooperation

and defection with respective frequencies x and 1 2 x.

The replicator dynamics are therefore given by a single

differential equation, _x ¼ x(PC � �P). With �P ¼ xPCþ(1� x)PD, we obtain

_x ¼ �x(1� x)(PD � PC): (2:2)

This differential equation has at least two equilibria: x ¼ 0,

at which all individuals defect, and x ¼ 1, at which all

individuals cooperate.

Our model extends the traditional public good game

[23–28] by incorporating incentives. Specifically, letting kdenote the number of cooperators among the n 2 1 co-

players in a group, the expected pay-offs for a defector and

a cooperator are given by

PD ¼Xn�1

k¼0

n� 1k

� �xk(1� x)n�1�k rck

n� b(1� w)nd

n� k

� �(2:3a)

and

PC ¼Xn�1

k¼0

n� 1k

� �xk(1� x)n�1�k rc(k þ 1)

n� cþ awnd

k þ 1

� �:

(2:3b)

Without incentives, d ¼ 0, we have PD 2 PC ¼ c(1 2 r/n) ¼ F,

which is the defector’s advantage in the public good game.

Page 3: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

rsif.royalsocietypublishing.orgJ.R.Soc.Interface

12:20140935

3

The replicator dynamics in equation (2.2) thus lead to full

defection, x ¼ 0, when r , n and to full cooperation, x ¼ 1,

when r . n. According to equations (2.3), incentives d . 0

modify the defector’s advantage for 0 , x , 1 as follows:

PD � PC

¼ F� ndXn�1

k¼0

n� 1

k

� �xk(1� x)n�1�k aw

k þ 1þ b(1� w)

n� k

� �:

(2:4a)

Usingn� 1

k

� �n

k þ 1¼ n

k þ 1

� �and

n� 1k

� �n

n� k¼ n

k

� �,

this yields

PD � PC ¼ F� d awXn�1

k¼0

nk þ 1

� �xk(1� x)n�1�k

"

þb(1� w)Xn�1

k¼0

nk

� �xk(1� x)n�1�k

#

¼ F� d awXn

k0¼1

nk0

� �xk0 (1� x)n�k0

x

"

þb(1� w)Xn�1

k¼0

nk

� �xk(1� x)n�k

1� x

#

¼ F� d aw1� (1� x)n

1� (1� x)þ b(1� w)

1� xn

1� x

� �

¼ F� d awXn�1

k¼0

(1� x)k þ b(1� w)Xn�1

k¼0

xk

" #:

(2:4b)

When rewarding or punishing are applied in isolation (w¼ 1

or w¼ 0), the defector’s advantage simplifies to F� adPn�1

k¼0

(1� x)k and F� bdPn�1

k¼0 xk, respectively [4,19]. Thus, the defec-

tor’s advantage is strictly increasing with the frequency x of

cooperators for pure rewarding and strictly decreasing with the

frequency x for pure punishing. In both the cases, there exists a

unique interior equilibrium of the replicator dynamics if and

only if the per capita incentive d lies within an intermediate

range, d2 , d , dþ, with

d� ¼Fan

and dþ ¼Fa: (2:5)

Here the value of a depends on the type of incentive being

applied: a¼ a for rewarding and a¼ b for punishing. The

unique interior equilibrium is globally asymptotically stable for

pure rewarding and unstable for pure punishing. Therefore,

when incentives are intermediate, d2 , d , dþ, the replicator

dynamics for pure rewarding lead to a mixture of defectors

and cooperators, while for pure punishing, they lead to bistability

between full defection and full cooperation. By contrast, if incen-

tives are very small, d� d2, or very large, d� dþ, the replicator

dynamics for pure rewarding and for pure punishing lead to

full defection or full cooperation, respectively.

3. ResultsWe first demonstrate, in §3.1, that an institutional sanctioning

policy we call ‘first carrot, then stick’, which switches from

rewarding to punishing when the frequency of cooperators

exceeds a threshold, minimizes the defector’s advantage. Its

effectiveness and efficiency are investigated in §3.2 and

compared with those of pure rewarding and pure punishing.

We extend our results to spatial populations in §3.3 and con-

clude by verifying, in §3.4, that our results are robust to other

parameter combinations and other model variants.

3.1. ‘First carrot, then stick’ as an optimal sanctioningpolicy

By allowing the fraction w of the incentive budget that is allo-

cated to rewarding, rather than to punishing, to change with

the frequency of cooperators, w ¼ w(x), we can represent a

broad range of institutional sanctioning policies. Below we

show that the ‘first carrot, then stick’ sanctioning policy is opti-

mal in that it minimizes the defector’s advantage PD 2 PC; it

thus maximizes the selection gradient _x at any frequency xof cooperators. This means that the ‘first carrot, then stick’

sanctioning policy results in the highest level of cooperation

for each parameter combination and that it consequently is

the most effective institutional sanctioning policy.

To see that the ‘first carrot, then stick’ sanctioning policy

minimizes the defector’s advantage, we first write equation

(2.4b) as

PD � PC ¼ F� d wXn�1

k¼0

(a(1� x)k � bxk)þ bXn�1

k¼0

xk

" #: (3:1)

As this equation is linear with respect to the weight w, a value

of w of either 0 or 1 is optimal depending on whether the sumPn�1k¼0 (a(1� x)k � bxk) is positive or negative, respectively. (In

the degenerate case when this sum equals zero, any value of

w will be optimal.) One can show that this sum is a decreas-

ing function of x with exactly one root x ¼ x satisfying

0 , x , 1, about which the sum changes sign from positive

to negative as x increases. Thus, PD 2 PC is minimized for

the following on–off control

w(x) ¼ 10

ifif

0 � x , x,x , x � 1,

�(3:2)

withPn�1

k¼0 (a(1�x)k � bxk) ¼ 0. This means that rewarding is

optimal when the fraction of cooperators is below x; other-

wise, punishing is optimal. For obvious reasons, we call

this institutional sanctioning policy ‘first carrot, then stick’.

3.2. Effectiveness and efficiency of ‘first carrot, thenstick’

Figure 1 shows how the replicator dynamics are affected by percapita incentives d under pure rewarding, pure punishing and

the optimal policy in equation (3.2), with the assumption that

rewarding is equally efficient (a ¼ b) or less efficient (a , b)

than punishing [32] (see also figure 2a–f). In particular, the

effects of the optimal policy, which are illustrated in figure

1e,f can be understood analytically. At the boundaries x ¼ 0

and x ¼ 1, PD 2 PC in equation (3.1) takes the values F 2 andand F 2 bnd, respectively. With the weight w(x) from equation

(3.2), PD 2 PC takes its maximum value, F� adPn�1

k¼0 (1�x)k

(or equivalently, F� bdPn�1

k¼0 x k), at x ¼ x; PD 2 PC is strictly

increasing for 0 � x , x and strictly decreasing for x , x � 1.

Hence, PD 2 PC¼ 0 can have at most two interior roots x.

Indeed, for d , minfF/(an),F/(bn)g, the replicator dynamics

converge to x¼ 0. For d . F/(an), a stable equilibrium enters

the interior state space 0 , x , 1 at x¼ 0, and the full-defection

equilibrium, x¼ 0, becomes unstable. For d . F/(bn), an

Page 4: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

0.5

1.0

0.2 0.4 0.6 0.8 1.0

0

per capita incentive, d

(a)

(d)

( f )

rew

ardi

ngpu

nish

ing

e

quili

briu

m f

requ

ency

of

coop

erat

ors

adap

tive

hybr

id

0.2 0.4 0.6 0.8 1.0

(c)

(e)

0.5

1.0

0

0.5

1.0

0 0

rewarding and punishing equally efficienta = b = 1

punishing more efficient than rewardinga = 1, b = 1.5(b)

Figure 1. Equilibrium cooperation frequencies of public good games in well-mixed populations for three institutional sanctioning policies. Each panel shows how thelocations of stable and unstable equilibrium cooperation frequencies (continuous and dashed lines, respectively) depend on the per capita incentive d. With no orvery small incentives d, full defection (x ¼ 0) is the only outcome, and for sufficiently large incentives d, so is full cooperation (x ¼ 1). This result applies to allthree considered sanctioning policies: pure rewarding with w ¼ 1, pure punishing with w ¼ 0, and the adaptive hybrid policy with w given by equation (3.2).Intermediate incentives d have strikingly different impacts, as follows. (a,b) Rewarding: When the institution increases the incentive beyond a threshold, a stableinterior equilibrium enters the state space at x ¼ 0, moves up to x ¼ 1, and eventually exits the state space at x ¼ 1. Consequently, full defection gives way tointermediate levels of cooperation, and finally to full cooperation. This outcome is independent of the initial condition. (c,d ) Punishing: When the institutionincreases the incentive beyond a threshold, an unstable interior equilibrium enters the state space at x ¼ 1, moves down to x ¼ 0, and eventually exits thestate space at x ¼ 0. Consequently, full defection gives way to bistability of full defection and full cooperation, and finally to full cooperation. In (d ), as theleverage b of punishing increases, full cooperation is established more readily, and the threshold value above which full cooperation is established regardlessof the initial condition is smaller than in (c). For ease of comparison, the thin dashed line in (d ) shows the unstable equilibria corresponding to (c). (e,f ) Adaptivehybrid: When the institution increases the incentive beyond a threshold, full defection gives rise to bistability of intermediate cooperation and full cooperation, andfinally to full cooperation. The two interior equilibria that arise, one stable and one unstable, annihilate each other for sufficiently large institutional incentives,leaving full cooperation as the global attractor. In ( f ), as the leverage b of punishing increases, full cooperation is established for smaller institutional incentives thanin (e). For ease of comparison, the thin continuous and dashed lines in ( f ) show the stable and unstable equilibria of (e), respectively. As punishing is more effectivethan rewarding, the switching threshold of the adaptive hybrid policy is below x ¼ 0.5. Parameters: n ¼ 5, r ¼ 2, c ¼ 1, a ¼ 1 and b ¼ 1 (a,c,e) or 1.5 (b,d,f ).

rsif.royalsocietypublishing.orgJ.R.Soc.Interface

12:20140935

4

unstable equilibrium enters the interior state space 0 , x , 1 at

x ¼ 1, and the full-cooperation equilibrium, x ¼ 1, becomes

stable. Thus, the interior state space 0 , x , 1 has both

stable and unstable equilibria if

maxFan

,Fbn

� �, d ,

F

aPn�1

k¼0 (1�x)k ¼F

bPn�1

k¼0 xk : (3:3)

As d increases within this interval, the two interior equilibria

approach each other and eventually merge and vanish at the

upper bound of this interval. For values of d beyond this

interval’s upper bound, the replicator dynamics converge to

x ¼ 1. In the special case when rewarding and punishing

are equally efficient, a ¼ b, the stable and unstable equilibria

enter the interior state space simultaneously (figure 1e). Con-

versely, in the extreme case when rewarding is much less

efficient than punishing, a� b, punishing alone is sufficient

from the very beginning, and hybridization with relatively

expensive rewarding is irrelevant.

For well-mixed populations, we have proved in §3.1 that the

hybridization of positive and negative incentives according to the

‘first carrot, then stick’ sanctioning policy described by equation

(3.2) minimizes the defector’s advantage, which ensures that this

sanctioning policy is most effective for converting a population of

Page 5: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

( f )

per capita incentive, d

4

3

2

1

5pu

blic

-ben

efit

fact

or, r

4

3

2

1

5(c)

4

3

2

1

5(e)

(b)

(d)

(l)

per capita incentive, d

(g)

(i)

(k)

0 0.2 0.4 0.6 0.8 1.00 0.2 0.4 0.6 0.8 1.00 0.2 0.4 0.6 0.8 1.00 0.2 0.4 0.6 0.8 1.0

(h)

(j)

rew

ardi

ngpu

nish

ing

adap

tive

hybr

id

establishment effectiveness recovery effectiveness establishment effectiveness recovery effectiveness

well-mixed populations spatial populations

initial condition:a single cooperator

initial condition:a single defector

initial condition:a single cooperator

initial condition:a single defector

0 0.2 0.4 0.6 0.8 1.0equilibrium frequency of cooperators

(a)

Figure 2. Effects of institutional sanctioning policies on public good games in well-mixed and spatial populations. The adaptive hybrid policy exhibits the broadestdomain of successfully establishing full cooperation (green) from an initially single cooperator ( first and third columns from the left), and also of recovering fullcooperation against an initially single defector (second and fourth columns from the left). With no or very small per capita incentives d, full defection (red) is theonly outcome, and with sufficiently large incentives, so is full cooperation; this result applies to all three institutional sanctioning policies. Intermediate incentiveshave strikingly different impacts, as follows. (a,b,g,h) Rewarding: In well-mixed populations, the outcome is independent of the initial condition; (a,b) are identical.In spatial populations, by contrast, full cooperation and full defection are more likely to be maintained when the public-benefit factor r is large and the per capitaincentive d is small, as seen in the upper left corners of (g) and (h), respectively. (c,d,i,j ) Punishing: When the institution increases d beyond a threshold value(which depends on r), full defection abruptly changes into full cooperation. The differences between (c) and (d ) or (i) and ( j ) indicate combinations of r and d forwhich full cooperation and full defection are both stable and for which the initial conditions, therefore, affect the outcome. The differences between (c) and (i)indicate that, interestingly, spatial population structure substantially reduces the range of combinations of r and d for which a single cooperator can invade,especially for large r. In (i)—and also in the upper parts of (g) and ( j ), as well as in the lower parts of (k) and (l )—the narrow ( yellow) band betweenno cooperation and full cooperation results from the survival probability of the initial cooperator (and therefore does not indicate the coexistence of cooperatorsand defectors). (e,f,k,l ) Adaptive hybrid: The domain of recovering full cooperation is almost equal to the case of punishing ( f,l), while the domain of establishingfull cooperation is much enlarged relative to the case of punishing, (e) and (k). In particular, as the institution increases d, the equilibrium frequency of cooperatorsgradually rises, and when d crosses a threshold value (again dependent on r), which is smaller than in the case of punishing, full cooperation is established abruptly(e,k). Parameters: n ¼ 5, c ¼ 1, a ¼ b ¼ 1, s ¼ 10 and N ¼ 100 (implying a population size of 10 000).

rsif.royalsocietypublishing.orgJ.R.Soc.Interface

12:20140935

5

defectors to cooperation (figure 2a–f). By combining the differen-

tial advantages of rewarding and punishing, this sanctioning

policy is far more effective than pure punishing in establishing

cooperation (figure 2c,e) and far more effective than pure reward-

ing in recovering cooperation (figure 2b,f ). Offering the ‘best of

both worlds’, the ‘first carrot, then stick’ sanctioning policy for

combining rewarding with punishing is therefore hereafter also

called the ‘adaptive hybrid’ sanctioning policy.

Although it is natural to expect that the threshold x at

which the adaptive hybrid policy switches from rewarding to

punishing might depend on other parameters, this is not the

case: the threshold remains the same independent of the percapita incentive d and the public-benefit factor r. What is more,

when there is no difference in leverage between positive and

negative incentives (a ¼ b), this threshold is situated at a 50% fre-

quency of cooperators, x ¼ 0:5. In practice, punishing is often

more effective than rewarding (a , b) [32], in which case the

switching point for optimal hybridization is situated at a fre-

quency of cooperators of less than 50%, x , 0:5 (figure 1f ).

The adaptive hybrid policy is not only more effective, but

also more efficient, for establishing and recovering cooperation

than either rewarding or punishing alone (figure 3a–f ).

Page 6: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

per capita incentive, d per capita incentive, d

4

3

2

1

5

0 0.2 0.4 0.6 0.8 1.0

4

3

2

1

5

4

3

2

1

5

0 0.2 0.4 0.6 0.8 1.0

rew

ardi

ngpu

nish

ing

publ

ic-b

enef

it fa

ctor

, rad

aptiv

e hy

brid

establishment costs recovery costs

cumulative costs

0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0

establishment costs recovery costs

cumulative costs

well-mixed populations spatial populations

104 105 106 103 1051070.1 10 105 104 106 108105 1030.1 10 107

initial condition:a single cooperator

initial condition:a single defector

initial condition:a single cooperator

initial condition:a single defector

( f )

(c)

(e)

(b)

(d)

(l)

(g)

(i)

(k)

(h)

(j)

(a)

Figure 3. Costs for establishing and recovering full cooperation. The adaptive hybrid policy is not only the most effective (figure 2) but also the least expensive inestablishing full cooperation from an initially single cooperator ( first and third columns from the left) and in recovering full cooperation against an initially singledefector (second and fourth columns from the left). If no or very small incentives are provided, achieving each of these goals is impossible (white), independent ofthe institutional policy. Otherwise, these policies have strikingly different impacts on the required cumulative costs. (a,b,g,h) Rewarding: Both in well-mixed and inspatial populations, rewarding requires recovery costs that are 1000 – 100 000 times more expensive than those for punishing or the adaptive hybrid policy. Fur-thermore, this relative cost difference increases in proportion to population size. (c,d ,i, j ) Punishing: In the case of punishing, recovery costs are much reduced relativeto the case of rewarding, while establishment costs remain at a similarly high level as, or are even slightly larger than, in the case of rewarding. (e,f,k,l ) Adaptivehybrid: The adaptive hybrid policy requires recovery costs that are similar to the case of punishing (and thus much lower than in the case of rewarding), butsubstantially reduces establishment costs relative to either rewarding or punishing. (For a detailed explanation of the costs at the border of the white regions,see electronic supplementary material, figure S1.) All parameters as in figure 2.

rsif.royalsocietypublishing.orgJ.R.Soc.Interface

12:20140935

6

Once a state of full cooperation has been reached, punishing is

cheaper as a means of recovering cooperation, as it needs to

be used only occasionally. As the adaptive hybrid policy stipu-

lates punishment once the frequency of cooperators surpasses

the threshold x, it is similar to pure punishment in this respect.

The two policies differ markedly, however, in the cost of

converting a population of defectors to a population of

cooperators. The adaptive hybrid policy has the lowest cumu-

lative cost of all three sanctioning policies, and hence requires

both the lowest establishment cost and the lowest recovery cost

for full cooperation. With respect to conversion speed, it gen-

erically takes a similar (finite) time for all three policies to

establish and recover cooperation (electronic supplementary

material, figure S1).

3.3. Extension to spatial populationsIn the real world, social planning tends to be spatially distribu-

ted and is often supported by sanctioning institutions. To see

whether the adaptive hybrid policy copes well with the result-

ing spatio-temporal complexity, we extend our framework to a

spatial population in which each individual inhabits one cell of

an N � N square lattice with periodic boundaries. In each gen-

eration, every individual on this lattice engages in a public good

game with its four nearest neighbours (n ¼ 5) and collects its

total pay-off through joining all the five games within its inter-

action neighbourhood. Then, the strategies of all individuals are

updated simultaneously. When a focal individual i’s strategy is

updated, a neighbour of it, j, is drawn at random among i’s four

nearest neighbours. Subsequently, individual i adopts its neigh-

bour j’s strategy with probability [1 þ exp(s(Ei 2 Ej))]21, where

Ei denotes individual i’s total pay-off and s determines the inten-

sity of selection [33,34]. The sanctioning institution receives

feedback from the five local participants, which means that

the x in equation (3.2) denotes the frequency of cooperators

within a given neighbourhood. The positive and negative incen-

tives determined by the adaptive hybrid policy therefore vary

across the lattice as local conditions require.

Page 7: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

t = 400

(b)

t = 600

t = 100t = 50 t = 200

t = 50

(c) adaptive hybrid

(a) punishingrewarding

t = 20 t = 100

t = 100generation t = 50 t = 200

cooperator defector

t = 200

booster stage

Figure 4. Emerging patterns of cooperation. For each incentive policy, the sequence of panels displays the spatio-temporal dynamics of cooperation, starting from asingle cooperator in a population of defectors. (a) Rewarding: A fragmented cluster of cooperators interspersed with defectors expands until small cooperator clustersoccur across the whole population (electronic supplementary material, movie S1). (b) Punishing: The initially single cooperator expands into a contiguous cluster ofcooperators, which eventually covers the entire population (electronic supplementary material, movie S2). (c) Adaptive hybrid: The initial spread of small cooperatorclusters closely resembles the case of rewarding. This development prepares the ground for local switches from rewarding to punishing, which drives the expansionof contiguous clusters of cooperators. This ‘booster stage’ enables the establishment of full cooperation for much lower incentives d than in the case of punishing(electronic supplementary material, movie S3). Parameters: r ¼ 2 and d ¼ 0.22 (a), 0.75 (b) or 0.22 (c). All other parameters as in figure 2.

rsif.royalsocietypublishing.orgJ.R.Soc.Interface

12:20140935

7

Also in spatial populations, the adaptive hybrid policy

turns out to be superior (figure 2g– l). Unexpectedly, it

gives rise to spatial patterns of cooperation and defection

that cannot easily be predicted from the patterns arising

under either rewarding or punishing alone. For small

and large incentives, the patterns emerging from the

adaptive hybrid policy when considering a single co-

operator in a population of defectors resemble the

patterns observed under pure rewarding and punishing,

respectively. Cooperators thrive under a policy of pure

rewarding (figure 4a), forming fragmented islands in

which they are locally interspersed with defectors, but ulti-

mately fail to establish full cooperation for the incentive

strength considered. Under a policy of pure punishing

(figure 4b), spatio-temporal dynamics are different: an inva-

sion that begins with a single cooperator in a population of

defectors always results in a contiguous cluster of coopera-

tors that grows and eventually displaces all defectors.

The adaptive hybrid policy, by contrast, for intermediate

incentive strengths exhibits an intriguing transition

between these two distinct patterns: fragmented islands of

cooperators, initially supported by rewarding, create cir-

cumstances under which punishing can act as a ‘booster

stage’ that capitalizes on and amplifies the pro-social effects

of rewarding, promoting the rapid growth of contiguous

clusters of cooperators that eventually displace all defectors

(figure 4c).

All three policies are capable of recovering cooperation in

much the same way as for well-mixed populations. The only

qualitative difference is that the successful spread of defec-

tors originating from an initially single defector can

occasionally split contiguous clusters of cooperators. This

phenomenon, however, which has previously been described

for the spatial extension of the well-studied Prisoner’s

Dilemma [35], occurs in our model only for vanishing or

very small incentives.

3.4. RobustnessIn the electronic supplementary material, we demonstrate the

robustness of our results with respect to the following model

variants. (i) First, we establish that in spatial populations the

adaptive hybrid policy with either local or global feedback

establishes and recovers full cooperation at lower cost and

under a wider range of conditions than an alternative hybrid-

ization of positive and negative incentives in which the

weight w is proportional to the frequency of cooperators,

w(x) ¼ x (electronic supplementary material, figure S2). We

also show that information about the local frequency of

cooperation allows a sanctioning institution that implements

the adaptive hybrid policy to establish full cooperation more

readily than information about the global (i.e. population-

wide) frequency of cooperation [20]. This result is in line

with expectations, as tailoring a strategy to local conditions

should generally achieve better results than a strategy that

depends on conditions averaged across large spatial scales.

We also explore (ii) a variant of the public good game in

which a cooperator does not benefit from its own contribution

[4,19] (electronic supplementary material, figure S3) and (iii) a

variant of the incentive scheme in which we relax the assumption

that the received incentive is inversely proportional to

the number of cooperators or defectors in an interacting

group [4,19] (electronic supplementary material, figure S4).

Furthermore, we test variants of our spatial model with

(iv) interactions encompassing the eight nearest neighbours

[33–35] (chess-king move, n ¼ 9, electronic supplementary

material, figure S5), (v) smaller population size (electronic sup-

plementary material, figure S6), (vi) asynchronous updating

Page 8: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

rsif.royalsocietypublishing.orgJ.R.Soc.Interface

12:2

8

[33,34] (electronic supplementary material, figure S7), (vii) a pro-

portional imitation rule [33,34] (electronic supplementary

material, figure S8), (viii) errors in perception and implemen-

tation (for individuals [36] or institutions [37], electronic

supplementary material, figures S9–S13) and (ix) varied switch-

ing thresholds (electronic supplementary material, figure S14).

We find that none of the variants (ii)–(viii) qualitatively affect

our results regarding the effectiveness and efficiency of the

three considered sanctioning policies (electronic supplementary

material, figures S3–S13). Exploring (ix) reveals that, for equal

leverages of positive and negative incentives, the optimal switch-

ing threshold for spatial populations approximately equals 50%,

just as in well-mixed populations (electronic supplementary

material, figures S14).

As a final model variant, we assume that (x) individuals

equally share the cost of funding the incentive budget (e.g.

through an entrance fee or poll tax) [4,19]. For this model var-

iant, we find the resulting dynamics to be entirely unaffected.

0140935

4. DiscussionHere we have demonstrated how an institutional sanctioning

policy of ‘first carrot, then stick’ can be surprisingly suc-

cessful in promoting cooperation. This policy establishes

and recovers cooperation at a lower cost and under a wider

range of conditions than either rewards or penalties alone

can do. Our findings are based on the public good game, a

standard framework for cooperation in groups. They apply

to both well-mixed and spatial populations and remain

robust under a broad spectrum of model variations and

parameter combinations.

To promote cooperation, rewards and penalties are fre-

quently used in concert. Considering how often they are

used together—at all levels of organization, from parents to

teachers to leaders of organizations—it is surprising that no

prior study to date has investigated the optimal use of a com-

bination of rewards and penalties in an institutional setting.

Here we have demonstrated that the optimal institutional

sanctioning policy is not given by a gradual change in the

relative allocation towards rewards and penalties, but by a

sudden switch from positive to negative incentives once

cooperation is sufficiently widespread. For example, teachers

who wish to establish order in an unruly group of pupils may

thus be well advised first to reward those pupils who behave

well and, once good behaviour has become the norm, shift to

reprimanding those pupils who do not. Similarly, leaders

faced with the difficult task of implementing an unpopular

directive may initially reward those who voluntarily

comply with it and, once the directive has been largely

phased in, make it mandatory, enforced by a penalty for

non-compliance. As these two examples indicate, the ‘first

carrot, then stick’ sanctioning policy can easily be adapted

to many real-life situations, making it a widely applicable

method for promoting cooperation.

Interestingly, when the ‘first carrot, then stick’ sanctioning

policy, here also called the adaptive hybrid policy, is used to

promote cooperation in spatial populations, it gives rise to

complex spatial patterns of cooperators and defectors that

differ qualitatively from the simpler patterns that arise

when rewards or penalties are used in isolation. This

is because punishing acts as a booster stage that reinforces

the pro-social effects of rewarding, thus allowing cooperation

to be rapidly established in those parts of a population

where cooperation has surpassed the critical threshold.

Although our analytical methods do not extend to spatial

populations, extensive numerical investigations confirm

that a sudden switch from rewarding to punishing, not a

gradual change in the relative allocation to these incentives,

also is optimal for promoting and recovering cooperation in

spatial populations.

Our theoretical results can be compared with the handful

of experimental studies that have explored the combined use

of positive and negative incentives in peer-to-peer sanction-

ing [38–41] or in sanctioning by an assigned team leader

[42]. Although these studies differ significantly in their exper-

imental design, they share two common characteristics. First,

punishment is typically more effective than rewarding at pro-

moting high contributions to the public good. Second,

players initially have a propensity for rewarding cooperation,

which is soon superseded by a propensity for punishing

defectors [38–40]. While the latter trend might superficially

be interpreted as corroborative evidence for the effectiveness

of the institutional sanctioning policy developed here, the

rationale for shifting from positive to negative incentives is

strikingly different. In the experimental studies, this shift

typically coincides with declining average contributions and

can thus be interpreted as a response to the emergence of

defectors [42]. In particular, a study on team leadership con-

cludes that ‘leaders who experience frequent complete free-

riding and high variance in contributions in their teams are

more likely to switch from positive to negative incentives’

[42], while other studies find that punishing is more effective

than rewarding at staving off complete free-riding [38–40].

By contrast, we have demonstrated the advantage of shifting

from positive to negative incentives as contributions increase,

and we predict that rewarding is more effective than punishing

in staving off complete free-riding [43].

We have determined the optimal sanctioning policy for a

social institution charged with overseeing rational agents.

Two complementary studies on peer-to-peer sanctioning that

account, respectively, for reputation effects and for the poten-

tial of group selection have similarly highlighted the role of

positive incentives in promoting incipient cooperation among

defectors [36,44]. These theoretical predictions, derived under

the assumption of rational behaviour, clearly question the

wisdom of the human behaviour observed in the aforemen-

tioned experimental studies. Understanding whether

punishment in the face of rampant defection is a human fal-

lacy or a rational choice under circumstances other than the

ones analysed here is a key challenge for future research.

Acknowledgements. We thank Karl Sigmund and Mitsuhiro Nakamurafor their valuable comments and suggestions.

Funding statement. This study was enabled by financial support by theAustrian Science Fund (FWF) to U.D. (TECT I-106 G11), through agrant for the research project The Adaptive Evolution of MutualisticInteractions as part of the multi-national collaborative research projectMutualisms, Contracts, Space, and Dispersal (BIOCONTRACT) selectedby the European Science Foundation as part of the European Colla-borative Research (EUROCORES) Programme The Evolution ofCooperation and Trading (TECT). U.D. gratefully acknowledgesadditional support by the European Commission, the EuropeanScience Foundation, the Austrian Ministry of Science and Research,and the Vienna Science and Technology Fund. T.S. acknowledgesadditional support by the Foundational Questions in EvolutionaryBiology Fund (RFP-12-21) and the Austrian Science Fund (FWF):P27018-G11.

Page 9: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

9

References

rsif.royalsocietypublishing.orgJ.R.Soc.Interface

12:20140935

1. Ostrom E. 1990 Governing the commons: theevolution of institutions for collective action.New York, NY: Cambridge University Press.

2. Sigmund K. 2007 Punish or perish? Retaliation andcollaboration among humans. Trends Ecol. Evol. 22,593 – 600. (doi:10.1016/j.tree.2007.06.012)

3. Balliet D, Mulder LB, Van Lange PAM. 2011 Reward,punishment, and cooperation: a meta-analysis.Psychol. Bull. 137, 594 – 615. (doi:10.1037/a0023489)

4. Sasaki T, Brannstrom A, Dieckmann U, Sigmund K.2012 The take-it-or-leave-it option allows smallpenalties to overcome social dilemmas. Proc. NatlAcad. Sci. USA 109, 1165 – 1169. (doi:10.1073/pnas.1115219109)

5. Oliver P. 1984 Rewards and punishment as selectiveincentives: an apex game. J. Conflict Resolut. 28,123 – 148. (doi:10.1177/0022002784028001007)

6. Boyd R, Richerson P. 1992 Punishment allows theevolution of cooperation (or anything else) insizable groups. Ethol. Sociobiol. 13, 171 – 195.(doi:10.1016/0162-3095(92)90032-Y)

7. Henrich J, Boyd R. 2001 Why people punishdefectors: weak conformist transmission canstabilize costly enforcement of norms in cooperativedilemmas. J. Theor. Biol. 208, 79 – 89. (doi:10.1006/jtbi.2000.2202)

8. Fehr E, Gachter S. 2002 Altruistic punishment inhumans. Nature 415, 137 – 140. (doi:10.1038/415137a)

9. Cuesta JA, Jimenez R, Lugo H, Sanchez A. 2008 Theshared reward dilemma. J. Theor. Biol. 251, 253 –263. (doi:10.1016/j.jtbi.2007.11.022)

10. Kanazawa T, Fukumoto Y, Ushio T, Misaka T. 2009Replicator dynamics with Pigovian subsidy andcapitation tax. Nonlinear Anal. Theory Methods Appl.71, e818 – e826. (doi:10.1016/j.na.2008.11.072)

11. Cressman R, Song J-W, Zhang B-Y, Tao Y. 2012Cooperation and evolutionary dynamics in thepublic goods game with institutional incentives.J. Theor. Biol. 299, 144 – 151. (doi:10.1016/j.jtbi.2011.07.030)

12. Cressman R, Wu J-J, Li C, Tao Y. 2013 Gameexperiments on cooperation through reward andpunishment. Biol. Theory 8, 158 – 166. (doi:10.1007/s13752-013-0106-2)

13. Falkinger J. 1996 Efficient private provision of publicgoods by rewarding deviations from average.J. Public. Econ. 62, 413 – 422. (doi:10.1016/0047-2727(95)01549-3)

14. Casari M, Luini L. 2009 Cooperation underalternative punishment institutions: an experiment.J. Econ. Behav. Organ. 71, 273 – 282. (doi:10.1016/j.jebo.2009.03.022)

15. Kosfeld M, Okada A, Riedl A. 2009 Institutionformation in public goods games. Am. Econ. Rev.99, 1335 – 1355. (doi:10.1257/aer.99.4.1335)

16. Sigmund K, De Silva H, Traulsen A, Hauert C. 2010Social learning promotes institutions for governingthe commons. Nature 466, 861 – 863. (doi:10.1038/nature09203)

17. Chen X, Gross T, Dieckmann U. 2013 Sharedrewarding overcomes defection traps in generalizedvolunteer’s dilemmas. J. Theor. Biol. 335, 13 – 21.(doi:10.1016/j.jtbi.2013.06.014)

18. Yang W, Liu W, Vina A, Tuanmu MN, He G, Dietz T,Liu J. 2013 Nonlinear effects of group size oncollective action and resource outcomes. Proc. NatlAcad. Sci. USA 110, 10 916 – 10 921. (doi:10.1073/pnas.1301733110)

19. Sasaki T. 2014 The evolution of cooperation throughinstitutional incentives and optional participation.Dyn. Games Appl. 4, 345 – 362. (doi:10.1007/s13235-013-0094-7)

20. Vasconcelos VV, Santos FC, Pacheco JM. 2013 Abottom-up institutional approach to cooperativegovernance of risky commons. Nat. Clim. Change 3,797 – 801. (doi:10.1038/nclimate1927)

21. Hilbe C, Traulsen A, Rohl T, Milinski M. 2014Democratic decisions establish stable authoritiesthat overcome the paradox of second-orderpunishment. Proc. Natl Acad. Sci. USA 111, 752 –756. (doi:10.1073/pnas.1315273111)

22. Abdallah S, Sayed R, Rahwan I, LeVeck BL, CebrianM, Rutherford A, Fowler JH. 2014 Corruption drivesthe emergence of civil society. J. R. Soc. Interface11, 20131044. (doi:10.1098/rsif.2013.1044)

23. Schelling TC. 1973 Hockey helmets, concealedweapons, and daylight saving: a study of binarychoices with externalities. J. Conflict Resol. 17,381 – 428. (doi:10.1177/002200277301700302)

24. Hamburger H. 1973 N-person Prisoner’s Dilemma.J. Math. Sociol. 3, 27 – 48. (doi:10.1080/0022250X.1973.9989822)

25. Dawes RM. 1980 Social dilemmas. Annul. Rev.Psychol. 31, 169 – 193. (doi:10.1146/annurev.ps.31.020180.001125)

26. Boyd R, Richerson PJ. 1988 The evolution ofreciprocity in sizable groups. J. Theor. Biol. 132,337 – 356. (doi:10.1016/S0022-5193(88)80219-4)

27. Hauert C, De Monte S, Hofbauer J, Sigmund K. 2002Replicator dynamics for optional public goodgames. J. Theor. Biol. 218, 187 – 194. (doi:10.1006/jtbi.2002.3067)

28. Hauert C, Michor F, Nowak MA, Doebeli M. 2006Synergy and discounting of cooperation in socialdilemmas. J. Theor. Biol. 239, 195 – 202. (doi:10.1016/j.jtbi.2005.08.040)

29. Hardin G. 1968 The tragedy of the commons.Science 162, 1243 – 1248. (doi:10.1126/science.162.3859.1243)

30. Herrmann B, Thoni C, Gachter S. 2008 Antisocialpunishment across societies. Science 319, 1362 –1367. (doi:10.1126/science.1153808)

31. Hofbauer J, Sigmund K. 1998 Evolutionary gamesand population dynamics. Cambridge, UK:Cambridge University Press.

32. Baron J. 2009 Weighing reward and punishment.Science 326, 1632. (doi:10.1126/science.326.5960.1632-a)

33. Roca CP, Cuesta JA, Sanchez A. 2009 Effect of spatialstructure on the evolution of cooperation. Phys. Rev.E 80, 046106. (doi:10.1103/PhysRevE.80.046106)

34. Szabo G, Fath G. 2007 Evolutionary games ongraphs. Phys. Rep. 446, 97 – 216. (doi:10.1016/j.physrep.2007.04.004)

35. Fu F, Nowak MA, Hauert C. 2010 Invasion andexpansion of cooperators in lattice populations:prisoner’s dilemma vs. snowdrift games. J. Theor.Biol. 266, 358 – 366. (doi:10.1016/j.jtbi.2010.06.042)

36. Hilbe C, Sigmund K. 2010 Incentives andopportunism: from the carrot to the stick.Proc. R. Soc. B 277, 2427 – 2433. (doi:10.1098/rspb.2010.0065)

37. Gachter S. 2012 Carrot or stick? Nature 483, 39 – 40.(doi:10.1038/483039a)

38. Sefton M, Shupp R, Walker JM. 2007 The effect ofrewards and sanctions in provision of public goods.Econ. Inquiry 45, 671 – 690. (doi:10.1111/j.1465-7295.2007.00051.x)

39. Kamijo Y, Takeuchi A. 2009 Voluntary contributionmechanism game and endogenous institutionselection. Waseda J. Polit. Sci. Econ. 368, 21 – 40 (inJapanese).

40. Sutter M, Haigner S, Kocher MG. 2010 Choosing thecarrot or the stick? Endogenous institutional choice insocial dilemma situations. Rev. Econ. Stud. 77,1540 – 1566. (doi:10.1111/j.1467-937X.2010.00608.x)

41. Andreoni J, Harbaugh WT, Vesterlund L. 2003 Thecarrot or the stick: rewards, punishments, andcooperation. Am. Econ. Rev. 93, 893 – 902. (doi:10.1257/000282803322157142)

42. Gurerk O, Irlenbusch B, Rockenbach B. 2009Motivating teammates: the leader’s choice betweenpositive and negative incentives. J. Econ. Psychol.30, 591 – 607. (doi:10.1016/j.joep.2009.04.004)

43. Szolnoki A, Perc M. 2012 Evolutionary advantages ofadaptive rewarding. New. J. Phys. 14, 093016.(doi:10.1088/1367-2630/14/9/093016)

44. Herold F. 2012 Carrot or stick? The evolution ofreciprocal preferences in a haystack model. Am. Econ.Rev. 102, 914 – 940. (doi:10.1257/aer.102.2.914)

Page 10: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 1 of 27

Electronic supplementary material for

First carrot, then stick:

How the adaptive hybridization of incentives promotes cooperation

Xiaojie Chen, Tatsuya Sasaki, Åke Brännström, and Ulf Dieckmann

Supplementary text and figures

In §S1, we present the individual-based model for our numerical investigations. In §S2, we investigate the

conversion time to full cooperation, i.e. how long it takes to establish or recover full cooperation (figure

S1). To examine the robustness of our main results, we finally, in §S3, consider a range of variants of our

main model (figures S2–S14).

S1. Individual-based model

In our individual-based model, the population is updated synchronously. In each generation, every indi-

vidual collects its total pay-off through joining a fixed number of games (in well-mixed populations, a

single game with 𝑛 − 1 randomly selected co-players, or in spatial populations, all possible games within

its interaction neighbourhood). Then, the strategies of all individuals are updated simultaneously. When

individual i’s strategy is updated, a neighbour j is drawn at random (in well-mixed populations, from the

entire population, or in spatial populations, among all of i’s 𝑛 − 1 neighbours). Subsequently, individual i

adopts its neighbour j’s strategy with probability [1 + exp(𝑠(𝐸! − 𝐸!))]!!, where 𝐸! denotes individual i’s

total pay-off and s determines the intensity of selection [33,34]. Cumulative costs in (finitely large) well-

mixed populations (figure 3) are calculated by means of individual-based model runs.

For spatial evolutionary games, it is well known that contiguous clusters of cooperators often fare

well in a population dominated by defectors, in which an isolated cooperator would achieve a pay-off well

below the population average [35]. Our numerical investigations respect the latter, most stringent, initial

condition. To study the conditions under which full cooperation can be stabilized, we further consider the

other extreme: a single defector in a population of cooperators. For infinite and well-mixed populations, ‘a

single cooperator/defector’ (e.g. in figure 2) means an infinitesimally small fraction of coopera-

tors/defectors. For finite spatial populations, the equilibrium frequency of cooperators is determined as the

fraction of cooperators in the whole population in the stationary state reached after sufficiently many gen-

erations. Each data point depicted in figures 2 and 3 represents the mean of 100 independent model runs.

Page 11: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 2 of 27

S2. Conversion to full cooperation: establishment and recovery

To better understand the factors that cause differences in cumulative costs between institutional sanction-

ing policies, we consider the conversion time to full cooperation. Specifically, by using the individual-

based model (§S1) we investigate the time required (i) to establish full cooperation from an initially single

cooperator, and (ii) to recover full cooperation after a single defector has entered the population. Although

it seems natural that these times would change significantly with parameters settings, it turns out that they

are largely independent of the public-benefit factor and the size of the institutional incentive, except when

these are reduced to values just above the threshold at which full cooperation can be established and re-

covered. In these cases, it can take an exceptionally long time to establish and recover cooperation, as seen

in figure S1.

Page 12: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 3 of 27

Figure S1. Establishment and recovery time. The establishment time describes how long it takes to achieve full cooperation from an initially single cooperator in a population of defectors. Similarly, the recovery time de-scribes how long it takes to resume full cooperation when a single defector is introduced into a population of cooperators. The times required to establish and recover cooperation are on the same order of magnitude for all three institutional sanctioning policies, except when the public-benefit factor and the per capita incentive are reduced to values just above the threshold at which full cooperation can be established and recovered. This is particularly visible for the recovery time in spatial populations (h). Note that rewarding is the most expensive policy for recovering cooperation. Parameters that are not varied are the same as in our main model (figure 2).

Page 13: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 4 of 27

S3. Alternative model variants

For spatial populations—and, in cases (ii) and (iii) below, also for well-mixed populations—we investi-

gate a range of model variants in which we change (i) the weight function and the feedback rule, (ii) the

specification of the public good game, (iii) the incentive scheme, (iv) the interaction neighbourhood, (v)

the population size, (vi) the population update scheme, (vii) the individual update rule, (viii) individual

and institutional errors of perception and implementation, and (ix) the switching threshold for the on-off

control. In particular, in (viii) we explicitly address a challenge highlighted in a recent review by Simon

Gächter: ‘real-life institutions do not work perfectly; for example, punishment or reward may not be cor-

rectly implemented. It remains unclear how these imperfectly applied incentives might affect cooperation’

[37].

(i) Weight function and feedback rule. We compare four variants of the spatial model in which

the feedback is either local, from the interaction neighbourhood, or global, from the whole population (as

in the case of well-mixed populations), and in which the weight function 𝑤 is either the on-off control or a

linear control under which 𝑤 is given by the frequency 𝑥 of cooperators. We find that, among those four

variants, the on-off control with local feedback is most widely successful, as well as least expensive, at

promoting full cooperation (figure S2).

(ii) Public good game. We consider a variant of the public good game, called ‘others-only’ [4,19].

Here, a player’s contribution is shared equally among the other 𝑛 − 1 co-players in an 𝑛-player group, so

that contributors have no direct gain from their own investments. In this variant, the public good game

constitutes a social dilemma, independent of the public-benefit factor 𝑟, because a player can always im-

prove her or his pay-off by withholding a contribution. In a group of 𝑛C cooperators and 𝑛D defectors

(with 𝑛C + 𝑛D = 𝑛 ), the pay-offs for a cooperator and a defector are 𝑟𝑐(𝑛C − 1)/(𝑛 − 1) − 𝑐 +

𝑎𝑤𝑛𝛿/𝑛C and 𝑟𝑐𝑛C/(𝑛 − 1) − 𝑏(1 − 𝑤)𝑛𝛿/𝑛D, respectively. Because there is no change in the last

terms of these expressions, describing the effects of incentives, equations (2.4), (2.5), (3.1), and (3.2) re-

main valid with the only change being that 𝐹 now equals 𝑐. In contrast to our main model, the others-only

variant has the property that 𝐹 is constant with respect to 𝑟, and thus, so are also the thresholds 𝛿! and 𝛿!.

Hence, the replicator dynamics in equation (2.2) are independent of 𝑟, and the equilibrium frequencies of

cooperators change only with respect to 𝛿 (figure S3−1a−f). Our main result, that the adaptive hybrid pol-

icy of using rewards and penalties is most efficient at establishing cooperation, holds also for the others-

only variant in well-mixed populations, as well as in spatial populations (figure S3−1g−l). In the latter

case, in striking contrast to our main model, larger values of 𝑟 make cooperation less likely to evolve

through punishing (figure S3−1i).

Page 14: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 5 of 27

(iii) Incentive scheme. We investigate a variant of the institutional sanctioning policy in which

the magnitude of incentives is not inversely proportional to the number of recipients [4]. This is intended

to mimic real-life situations in which the expected incentive is constant or near-constant as long as the

number of recipients is not unmanageably large. Following earlier work on positive and negative incen-

tives [4], we generalize the expected penalty as 𝑏 1 − 𝑤 𝑛𝛿/[ 1 − ℎ + ℎ(𝑛 − 𝑛C)], and the expected

reward as 𝑎𝑤𝑛𝛿/[(1 − ℎ) + ℎ𝑛C]. In these expressions, the parameter ℎ controls how the magnitude of

incentives changes with the number of recipients, 0 ≤ ℎ ≤ 1. When ℎ = 0, incentives are independent of

the number of recipients, whereas for increasing values of ℎ, incentives become less uniform. Our main

model corresponds to the case ℎ = 1. Our result is qualitatively unchanged for ℎ = 0.8, 0.6, 0.4, or 0.2 in

the establishment case, and for ℎ = 0.8 or 0.6 in the recovery case. For much smaller values of ℎ, the ad-

vantage of the adaptive hybrid policy is lessened, and pure rewarding and pure punishing are similarly

effective at achieving a high level of cooperation for a broad range of parameters (figures S4–1 and S4–3).

The adaptive hybrid policy, however, still remains least expensive, except on the verge of full cooperation

(figures S4–2 and S4–4).

(iv) Interaction neighbourhood. We investigate a variant for spatial populations in which inter-

actions occur in the Moore neighbourhood [33,34], so a focal individual interacts with individuals in the

eight nearest cells (which might be reached by one move of a chess king). Individual-based model runs

confirm that our main result is qualitatively unaffected by this enlarged neighbourhood (figure S5).

(v) Population size. We reduce the size of the square lattice on which individual-based model

runs take place from 100×100 to 10×10, and find that our main result is qualitatively robust under such

downsizing (figure S6).

(vi) Population update scheme. As a further variant for spatial populations, we consider asyn-

chronous updating [33,34], in which in each time step one individual is chosen at random from the popula-

tion, and immediately updated. Our main result, derived for synchronous updating (§S1), is robust under

such asynchronous updating (figure S7).

(vii) Individual update rule. We also consider a variant of the individual update rule, in which a

focal individual 𝑖 with total pay-off 𝐸!  adopts the strategy of an individual 𝑗 with total pay-off 𝐸!, to whom

the focal individual is being compared, with a probability proportional to their pay-off difference [33,34],

provided 𝐸! − 𝐸! > 0. Specifically, the imitation probability is given by

𝜃!→! =(𝐸! − 𝐸!)/∆          if  𝐸! − 𝐸! > 0,              0                                  otherwise,                                                                                                                  (S1)  

in which the scaling factor ∆ ensures that 0 ≤ 𝜃!→! ≤ 1. Our main result remains robust under this varia-

tion (figure S8), and changing the value of ∆ does not affect the equilibrium frequency of cooperators

(instead, it only scales the resulting cumulative cost).

Page 15: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 6 of 27

(viii) Individual and institutional errors. We investigate the effects on our main result of five

different types of possible errors in perception and implementation for individuals [36] and institutions

[37]. None of the errors considered qualitatively affects our findings:

• Strategy-implementation error (figure S9). When an individual participates in the public

goods game, it uses the strategy opposite to its own with probability 𝑢!.

• Strategy-observation error (figure S10). To gauge incentives, the institution must observe

each individual’s strategy. We assume that, with probability 𝑢!, the institution incorrectly ob-

serves an individual’s strategy.

• Incentive-distribution error (figure S11). After the public good game, the institution applies

incentives to the target individuals in a group. We assume that each of them correctly receives

their respective incentive independently with probability 1 − 𝑢!; otherwise, the incentive is

randomly given to another individual in the group. Consequently, a cooperator (defector) may

be punished (rewarded), and multiple instances of rewarding and/or punishing may occur for a

single individual.

• Payoff-observation error (figure S12). Each individual reviews its strategy by using a stochas-

tic update rule following the so-called Fermi function, which depends on the pay-off differ-

ence between the focal individual and a randomly selected neighbour (§S1). In this function,

we may reduce the strength 𝑠 of selection to represent an increasing degree of erroneous per-

ception of a neighbour’s total pay-off.

• Strategy-imitation error (figure S13). Depending on the individual update rule, each individu-

al makes a decision on whether or not to imitate a neighbour’s strategy. If an individual de-

cides to do so, we assume that it mistakenly imitates the wrong strategy with probability 𝑢!.

In contrast to the four aforementioned sources of errors, we find that such strategy-imitation

errors facilitate the establishment of cooperation from an initially single cooperator in a popu-

lation of defectors, which becomes successful for larger values of the public-benefit factor 𝑟,

in particular, in the case of punishing (figure S13–1c).

(ix) Switching threshold. Finally, we explore which switching threshold for the on-off control

most efficiently promotes cooperation in the spatial model. We address this question for interactions with

the four nearest neighbours (figures S14–1 and S14–2), as well as for interactions with the eight nearest

neighbours (figures S14–3 and S14–4). In both cases, the institution switches from rewarding to punishing

when the number of cooperators in a group exceeds a given threshold 𝑛. We find that for 𝑛 = 2 in the

former case (𝑛 = 5) and for 𝑛 around 4 in the latter case (𝑛 = 9), the spatial on-off control most effective-

ly establishes full cooperation from an initially single cooperator. This result closely matches predictions

of the analytical theory, 𝑛 = 2.5 and 𝑛 = 4.5, respectively.

Page 16: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 7 of 27

Figure S2-1. Effects of institutional sanctioning policies on spatial public good games under different feed-backs and hybridizations of incentives. The on-off control based on local feedback (e) is most effective for establishing and recovering cooperation. This control and feedback are used by the adaptive hybrid policy. Parameters that are not varied are the same as in figure 2.

Page 17: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 8 of 27

Figure S2-2. Costs for establishing and recovering full cooperation in spatial public good games under differ-ent feedbacks and hybridizations of incentives. The on-off control based on local feedback (e) is the least ex-pensive option for establishing and recovering cooperation. This control and feedback are used by the adaptive hybrid policy. Parameters that are not varied are the same as in figure 2.

Page 18: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 9 of 27

Figure S3-1. Effects of institutional sanctioning policies on ‘others-only’ public good games. The adaptive hybrid policy has the broadest domain of success also when individuals do not receive any direct benefits from their own investments. Note that in (i) the line that separates the parameter regions for full defection (red) and full cooperation (green) has positive slope, whereas, in our main model this boundary has negative slope (fig-ure 2i). Parameters that are not varied are the same as in figure 2.

Page 19: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 10 of 27

Figure S3-2. Costs for establishing and recovering full cooperation in ‘others-only’ public good games. The adaptive hybrid policy is the least expensive also when individuals do not receive any direct benefits from their own investments. Parameters that are not varied are as in figure 2.

Page 20: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 11 of 27

 

Figure S4-1. Effects of institutional sanctioning policies on spatial public good games when the magnitudes of incentives are uniform. For decreasing values of ℎ, the magnitudes of incentives become more uniform. The adaptive hybrid policy has the broadest domain of success in establishing cooperation also when the magni-tudes of incentives are not inversely proportional to the number of recipients in a focal individual’s interaction group. Parameters that are not varied are the same as in figure 2.    

Page 21: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 12 of 27

 Figure S4-2. Costs for establishing full cooperation in spatial public good games when the magnitudes of in-centives are uniform. For decreasing values of ℎ, the magnitudes of incentives become more uniform. The adaptive hybrid policy remains least expensive also when the magnitudes of incentives are not inversely pro-portional to the number of recipients in a focal individual’s interaction group. The only minor exceptions occur at the verge of full cooperation. Parameters that are not varied are the same as in figure 2.    

Page 22: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 13 of 27

 

Figure S4-3. Effects of institutional sanctioning policies on spatial public good games when the magnitudes of incentives are uniform. For decreasing values of ℎ, the magnitudes of incentives become more uniform. The adaptive hybrid policy has the broadest domain of success in establishing cooperation also when the magni-tudes of incentives are not inversely proportional to the number of recipients in a focal individual’s interaction group. Parameters that are not varied are the same as in figure 2.

Page 23: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 14 of 27

Figure S4-4. Costs for recovering full cooperation in spatial public good games when incentives are uniform in size. For decreasing values of ℎ, the magnitudes of incentives become more uniform. The adaptive hybrid poli-cy remains least expensive also when the magnitudes of incentives are not inversely proportional to the number of recipients in a focal individual’s interaction group. The only minor exceptions occur at the verge of full co-operation. Parameters that are not varied are the same as in figure 2.

Page 24: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 15 of 27

Figure S5-1. Effects of institutional sanctioning poli-cies on spatial public good games with enlarged inter-action neighbourhood. The adaptive hybrid policy has the broadest domain of success also when individuals interact with their eight nearest neighbours (𝑛 = 9). Parameters that are not varied are the same as in figure 2.

Figure S5-2. Costs for establishing and recovering full cooperation in spatial public good games with en-larged interaction neighbourhood. The adaptive hybrid policy is the least expensive also when individuals interact with their eight nearest neighbours (𝑛 = 9). Parameters that are not varied are the same as in figure 2.

Page 25: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 16 of 27

Figure S6-1. Effects of institutional sanctioning poli-cies on spatial public good games with smaller popula-tion size. The adaptive hybrid policy has the broadest domain of success also for populations on a 10×10 periodic square lattice. Parameters that are not varied are the same as in figure 2.

Figure S6-2. Costs for establishing and recovering full cooperation in spatial public good games with smaller population size. The adaptive hybrid policy is the least expensive also for populations on a 10×10 periodic square lattice. Parameters that are not varied are the same as in figure 2.

Page 26: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 17 of 27

Figure S7-1. Effects of institutional sanctioning poli-cies on spatial public good games with asynchronous updating. The adaptive hybrid policy has the broadest domain of success also when individual strategies are sequentially reviewed and updated. Parameters that are not varied are the same as in figure 2.

Figure S7-2. Costs for establishing and recovering full cooperation in spatial public good games with asyn-chronous updating. The adaptive hybrid policy is the least expensive also when individual strategies are sequentially reviewed and updated. Parameters that are not varied are the same as in figure 2.

Page 27: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 18 of 27

Figure S8-1. Effects of institutional sanctioning poli-cies on spatial public good games with the proportion-al imitation rule. The adaptive hybrid policy has the broadest domain of success also when decisions on updating an individual’s strategy follow the propor-tional imitation rule, in which the scaling factor ∆ in equation (S1) is given by 𝑛[𝑐(𝑟 + 1) + 2𝑛𝛿]. Parame-ters that are not varied are the same as in figure 2.

Figure S8-2. Costs for establishing and recovering full cooperation in spatial public good games with the proportional imitation rule. The adaptive hybrid policy is the least expensive also when decisions on updating an individual’s strategy follow the proportional imita-tion rule. Parameters that are not varied are the same as in figure 2.

Page 28: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 19 of 27

Figure S9-1. Effects of institutional sanctioning poli-cies on spatial public good games with strategy-implementation errors. The adaptive hybrid policy has the broadest domain of success also when individuals participating in the public good game mistakenly use the strategy opposite to their own with probability 𝑢! = 0.01. Parameters that are not varied are the same as in figure 2.

Figure S9-2. Costs for establishing and recovering full cooperation in spatial public good games with strate-gy-implementation errors. The adaptive hybrid policy is the least expensive also when individuals participat-ing in the public good game mistakenly use the strate-gy opposite to their own with probability 𝑢! = 0.01. Parameters that are not varied are the same as in figure 2.

Page 29: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 20 of 27

Figure S10-1. Effects of institutional sanctioning poli-cies on spatial public good games with strategy-observation errors. The adaptive hybrid policy has the broadest domain of success also when the sanctioning institution mistakenly observes the opposite of an individual’s strategy with probability 𝑢! = 0.01. Pa-rameters that are not varied are the same as in figure 2.

Figure S10-2. Costs for establishing and recovering full cooperation in spatial public good games with strategy-observation errors. The adaptive hybrid policy is the least expensive also when the sanctioning insti-tution mistakenly observes the opposite of an individ-ual’s strategy with probability 𝑢! = 0.01. Parameters that are not varied are the same as in figure 2.

Page 30: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 21 of 27

Figure S11-1. Effects of institutional sanctioning poli-cies on spatial public good games with incentive-distribution errors. The adaptive hybrid policy has the broadest domain of success also when the institution, in the course of distributing incentives to individuals in an interaction group, mistakenly applies an incen-tive intended for an individual to another individual randomly selected from the group with probability 𝑢! = 0.01. Parameters that are not varied are the same as in figure 2.

Figure S11-2. Costs for establishing and recovering full cooperation policies in spatial public good games with incentive-distribution errors. The adaptive hybrid policy is the least expensive also when the institution, in the course of distributing incentives to individuals in an interaction group, mistakenly applies an incen-tive intended for an individual to another individual randomly selected from the group with probability 𝑢! = 0.01. Parameters that are not varied are the same as in figure 2.

Page 31: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 22 of 27

Figure S12-1. Effects of institutional sanctioning policies on spatial public good games with different payoff-observation errors. The adaptive hybrid policy has the broadest domain of success also when the ability of individuals to correctly update their strategy is varied. The Fermi function in the individual updating rule (§S1) is considered for two different values of the strength 𝑠 of selection. The two columns on the left and on the right, respectively, corre-spond to stronger selection (𝑠 = 100), and thus less payoff-observation errors, and to weaker selection (𝑠 = 1), and thus more payoff-observation errors, than in our main model (𝑠 = 10 in figure 2). Parameters that are not varied are the same as in figure 2.

Page 32: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 23 of 27

Figure S12-2. Costs for establishing and recovering full cooperation policies in spatial public good games with vary-ing payoff-observation errors. The adaptive hybrid policy is the least expensive also when the ability of individuals to correctly update their strategy is varied. The Fermi function in the individual updating rule (§S1) is considered for varying degree of the strength 𝑠 of selection. The two columns on the left and on the right correspond, respectively, to stronger selection (𝑠 = 100), and thus less payoff-observation errors, and to weaker selection (𝑠 = 1), and thus more payoff-observation errors, than in our main model (𝑠 = 10 in figure 2). Parameters that are not varied are the same as in figure 2.

Page 33: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 24 of 27

Figure S13-1. Effects of institutional sanctioning poli-cies on spatial public good games with strategy-imitation errors. The adaptive hybrid policy has the broadest domain of success also if, when imitating a neighbour’s strategy, the focal individual mistakenly chooses the opposite strategy with probability 𝑢! = 0.01. Parameters that are not varied are the same as in figure 2.

Figure S13-2. Costs for establishing and recovering full cooperation in spatial public good games with strategy-imitation errors. The adaptive hybrid policy is the least expensive also if, when imitating a neigh-bour’s strategy, the focal individual mistakenly choos-es the opposite strategy with probability 𝑢! = 0.01. As the equilibrium state of this model variant always includes some defectors, we consider cooperation to be established when the fraction of cooperators ex-ceeds a threshold of 99%. For recovery costs, we con-sider the average cost over 1,000 time steps. The white regions in the right column show parameter combina-tions for which the fraction of cooperators fell below this threshold and did not recover within the given time frame. Results are qualitatively unchanged for other thresholds and time frames. Parameters that are not varied are the same as in figure 2.

Page 34: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 25 of 27

Figure S14-1. Effects on spatial public good games of institutional sanctioning policies based on on-off con-trols with varying switching thresholds. Switching from rewarding to punishing when the number of co-operators in a group exceeds 𝑛 = 2 is the most effec-tive on-off control for establishing full cooperation. This can be compared with the theoretical prediction of optimal switching at 50% cooperators in well-mixed populations, implying 𝑛 = 2.5 for an interac-tion neighbourhood of five individuals (𝑛 = 5). Pa-rameters that are not varied are the same as in figure 2.

Figure S14-2. Costs for establishing and recovering full cooperation in spatial public good games for on-off controls with different switching thresholds. Switching from rewarding to punishing when the number of cooperators in a group exceeds 𝑛 = 2 is the least expensive on-off control for establishing full cooperation. Parameters that are not varied are the same as in figure 2.

Page 35: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 26 of 27

Figure S14-3. Effects on spatial public good games with an enlarged interaction neighbourhood of institutional sanc-tioning policies based on on-off controls with varying switching thresholds. Switching from rewarding to punishing when the number of cooperators in a group exceeds 𝑛 = 4 is the most effective on-off control for establishing full cooperation when the interaction neighbourhood consists of nine individuals (𝑛 = 9). This can be compared with the theoretical prediction of optimal switching at 50% cooperators in well-mixed populations, implying 𝑛 = 4.5 for an interaction neighbourhood of nine individuals. Parameters that are not varied are the same as in figure 2.

Page 36: First carrot, then stick: how the adaptive hybridization of …dieckman/reprints/ChenEtal2015.pdf · exceeds the pay-off difference between a cooperator and a defector, regardless

Page 27 of 27

Figure S14-4. Costs for establishing and recovering full cooperation in spatial public good games with an enlarged interaction neighbourhood for on-off controls with different switching thresholds. Switching from rewarding to pun-ishing when the number of cooperators in a group exceeds 𝑛 = 4 is the least expensive on-off control for establishing full cooperation when the interaction neighbourhood consists of nine individuals (𝑛 = 9). Parameters that are not varied are the same as in figure 2.


Recommended