web appendices to selling to overcon–dent consumers

22
Web Appendices to Selling to Overcondent Consumers Michael D. Grubb MIT Sloan School of Management Cambridge, MA 02142 [email protected] www.mit.edu/~mgrubb May 2, 2008 B Option Pricing Intuition This appendix provides additional intuition based on option pricing for the result in Proposition 2. Consider the case of monopoly. At time one, the monopolist is selling a series of call options, or equivalently units bundled with put options, rather than units themselves. The marginal price charged for a unit q at time two is simply the strike price of the option sold on unit q at time one. The series of call options being sold are interrelated; a call option for unit q cant be exercised unless the call option for unit q 1 has already been exercised. However, it is useful to consider the market for each option independently. According to Proposition 2 (equation 8), when there is no pooling and the satiation constraint is not binding, the optimal marginal price for a unit q is: ^ P q (q)= C q (q)+ V q (q; ^ (q)) F ( ^ (q)) F ( ^ (q)) f ( ^ (q)) (21) It turns out that this is exactly the strike price that maximizes the net value of a call or put option on unit q given the di/erence in priors between the two parties. 37 To show this explicitly, write the net value NV of a call option on minute q as the di/erence between the consumersvalue of the option CV and the rms cost of providing that option, FV . The option will be exercised whenever the consumer values unit q more than the strike price p, that is whenever V q (q;) p. Let (p) denote the minimum type who exercises the call option, characterized by the equality V q (q; (p)) = p. 37 This parallels Mussa and Rosens (1978) nding in their static screening model, that the optimal marginal price for unit q is identical to the optimal monopoly price for unit q if the market for unit q were treated independently of all other units. 1

Upload: others

Post on 11-Feb-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Web Appendices to Selling to Overcon–dent Consumers

Web Appendices toSelling to Overcon�dent Consumers

Michael D. Grubb

MIT Sloan School of Management

Cambridge, MA 02142

[email protected]

www.mit.edu/~mgrubb

May 2, 2008

B Option Pricing Intuition

This appendix provides additional intuition based on option pricing for the result in Proposition 2.

Consider the case of monopoly. At time one, the monopolist is selling a series of call options,

or equivalently units bundled with put options, rather than units themselves. The marginal price

charged for a unit q at time two is simply the strike price of the option sold on unit q at time one.

The series of call options being sold are interrelated; a call option for unit q can�t be exercised

unless the call option for unit q � 1 has already been exercised. However, it is useful to considerthe market for each option independently. According to Proposition 2 (equation 8), when there is

no pooling and the satiation constraint is not binding, the optimal marginal price for a unit q is:

P̂q (q) = Cq (q) + Vq�(q; �̂ (q))F �(�̂ (q))� F (�̂ (q))

f(�̂ (q))(21)

It turns out that this is exactly the strike price that maximizes the net value of a call or put option

on unit q given the di¤erence in priors between the two parties.37

To show this explicitly, write the net value NV of a call option on minute q as the di¤erence

between the consumers�value of the option CV and the �rm�s cost of providing that option, FV .

The option will be exercised whenever the consumer values unit q more than the strike price p,

that is whenever Vq (q; �) � p. Let � (p) denote the minimum type who exercises the call option,

characterized by the equality Vq (q; � (p)) = p.

37This parallels Mussa and Rosen�s (1978) �nding in their static screening model, that the optimal marginal pricefor unit q is identical to the optimal monopoly price for unit q if the market for unit q were treated independently ofall other units.

1

Page 2: Web Appendices to Selling to Overcon–dent Consumers

The consumers� value for the option is their expected value received upon exercise, less the

expected strike price paid, where expectations are based on the consumers�prior F � (�):

CV (p) =

Z ��

�(p)Vq (q; �) f

� (�) d� � [1� F � (� (p))] p

The �rm�s cost of providing the option is the probability of exercise based on the �rm�s prior F (�)

times the di¤erence between the cost of unit q and the strike price received:

FV (p) = [1� F (� (p))] (c� p)

Putting these two pieces together, the net value of the call option is equal to the consumers�

expected value of consumption less the �rms expected cost of production plus an additional term

due to the gap in perceptions:

NV (p) =

Z ��

�(p)Vq (q; �) f

� (�) d� � [1� F (� (p))] c+ [F � (� (p))� F (� (p))] p

The additional term [F � (� (p))� F (� (p))] p represents the di¤erence between the exercise paymentthe �rm expects to receive and the consumer expects to pay. The term [F � (� (p))� F (� (p))]represents the disagreement between the parties about the probability of exercise.

As a monopolist selling call options on unit q earns the net value NV (p) of the call option

by charging the consumer CV (p) upfront, a monopolist should set the strike price p to maximize

NV (p). By the implicit function theorem, ddp� (p) =

1Vq�(q;�(p))

, so the �rst order condition which

characterizes the optimal strike price is:

f (� (p))

Vq� (q; � (p))[p� Cq (q)] = [F � (� (p))� F (� (p))] (22)

As claimed earlier, this is identical to the characterization of the optimal marginal price P̂q (q)

for the complete nonlinear pricing problem when monotonicity and satiation constraints are not

binding (equation 21).

Showing that the optimal marginal price for unit q is given by the optimal strike price for a

call option on unit q is useful, because the �rst order condition q (q; �) = 0 can be interpreted

in the option pricing framework. Consider the choice of exercise price p for an option on unit q.

A small change in the exercise price has two e¤ects. First, if a consumer is on the margin, it will

change the consumers�exercise decision. Second, it changes the payment made upon exercise by all

infra-marginal consumers. In a common-prior model, the infra-marginal e¤ect would net to zero as

2

Page 3: Web Appendices to Selling to Overcon–dent Consumers

the payment is a transfer between the two parties. This is not the case here, however, as the two

parties disagree on the likelihood of exercise by [F � (� (p))� F (� (p))].Consider the �rst order condition as given above in equation (22). On the left hand side, the

term f(�(p))Vq�(q;�(p))

represents the probability that the consumer is on the margin and that a marginal

increase in the strike price p would stop the consumer exercising. The term [p� Cq (q)] is the costto the �rm if the consumer is on the margin and no longer exercises. There is no change in the

consumer�s value of the option by a change in exercise behavior at the margin, as the margin is

precisely where the consumer is indi¤erent to exercise (Vq (q; � (p)) = p).

On the right hand side, the term [F � (� (p))� F (� (p))] is the �rm�s gain on infra-marginalconsumers from charging a slightly higher exercise price. This is because consumers believe they

will pay [1� F � (� (p))] more in exercise fees, and therefore are willing to pay [1� F � (� (p))] lessupfront for the option. However the �rm believes they will actually pay [1� F (� (p))] more inexercise fees, and the di¤erence [F � (� (p))� F (� (p))] is the �rm�s perceived gain.

The �rst order condition requires that at the optimal strike price p, the cost of losing mar-

ginal consumers f(�(p))Vq�(q;�(p))

[p� Cq (q)] is exactly o¤set by the "perception arbitrage" gain on infra-marginal consumers [F � (� (p))� F (� (p))].

Setting the strike price above or below marginal cost is always costly because it reduces e¢ ciency.

In the discussion above, referring to [F � (� (p))� F (� (p))] as a "gain" to the �rm for a marginal

increase in strike price implies that the term [F � (� (p))� F (� (p))] is positive. This is the casefor � (p) > ��, when consumers underestimate their probability of exercise. In this case, from the

�rm�s perspective, raising the strike price above marginal cost increases pro�ts on infra-marginal

consumers, thereby e¤ectively exploiting the perception gap. On the other hand, for � (p) < ��, the

term [F � (� (p))� F (� (p))] is negative and consumers overestimate their probability of exercise.In this case reducing the strike price below marginal cost exploits the perception gap between

consumers and the �rm.

Fixing � and the �rm�s prior F (�), the absolute value of the perception gap is largest when the

consumer�s prior is at either of two extremes, F � (�) = 1 or F � (�) = 0. When F � (�) = 1, the

optimal marginal price reduces to the monopoly price for unit q where the market for minute q is

independent of all other units. This is because consumers believe there is zero probability that they

will want to exercise a call option for unit q. The �rm cannot charge anything for an option at time

one; essentially the �rm must wait to charge the monopoly price until time two when consumers

realize their true value.

Similarly, when F � (�) = 0, the optimal marginal price reduces to the monopsony price for unit

q. Now rather than thinking of a call option, think of the monopolist as selling a bundled unit and

3

Page 4: Web Appendices to Selling to Overcon–dent Consumers

put option at time one. In this case consumers believe they will consume the unit for sure and

exercise the put option with zero probability. This means that the �rm cannot charge anything for

the put option upfront, and must wait until time two when consumers learn their true values and

buy units back from them at the monopsony price. The �rm�s ability to do so is of course limited

by free disposal which means the �rm could not buy back units for a negative price.

Marginal price can therefore be compared to three benchmarks. For all quantities q, the marginal

price will lie somewhere between the monopoly price pml (q) and the maximum of the monopsony

price pms (q) and zero, hitting either extreme when F � (�) = 1 or F � (�) = 0, respectively. When

F � (�) = F (�), marginal price is equal to marginal cost. To illustrate this point, the equilibrium

marginal price for the running example with positive marginal cost c = $0:035 and low overcon�-

dence� = 0:25 previously shown in Figure 4, plot C is replotted with the monopoly and monopsony

prices for comparison in Figure 10.

300 400 500 600 700­0.1

0

0.1

0.2

0.3

quantity

$

Optimal MPMonopolyMonopsonyMC

Figure 10: Equilibrium pricing for Example 1 given c = $0:035 and � = 0:25: Marginal price isplotted along with benchmarks: (1) marginal cost, (2) ex post monopoly price - the upper bound,and (3) ex post monopsony price - the lower bound.

C Pooling

As it was omitted from the main text, a characterization of pooling quantities when the monotonic-

ity constraint binds is provided below in Lemma 4. This is useful because it facilitates the calcu-

lation of pooling quantities in numerical examples.

Lemma 4 On any interval [�1; �2] such that monotonicity (but not non-negativity) is binding in-

side, but not just outside the interval, the equilibrium allocation is constant at some level q̂ > 0 for

all � 2 [�1; �2]. Further, the pooling quantity q̂ and bounds of the pooling interval [�1; �2] must main-

4

Page 5: Web Appendices to Selling to Overcon–dent Consumers

tain continuity of q̂ (�) and satisfy the �rst order condition on average:R �2�1q (q̂; �) f (�) d� = 0.

Non-negativity binds inside, but not just above, the interval [�; �2] only ifR �2� q (0; �) f (�) d� � 0

and qR (�2) = 0.

Proof. Given the result in Lemma 2, the proof is omitted as it closely follows ironing results

for the standard screening model. The result follows from the application of standard results in

optimal control theory (Seierstad and Sydsæter 1987). See for example the analogous proof given

in Fudenberg and Tirole (1991), appendix to chapter 7.

Note that because the virtual surplus function is strictly quasi-concave, but not necessarily

strictly concave, optimal control results yield necessary rather than su¢ cient conditions for the

optimal allocation. For the special case Vqq� = 0, virtual surplus is strictly-concave and applying

the ironing algorithm suggested by Fudenberg and Tirole (1991) using the conditions in Lemmas 2

and 4 is su¢ cient to identify the uniquely optimal allocation.

Proposition 2 characterizes marginal pricing at quantities for which there is no pooling, and

states that marginal price will jump discretely upwards at quantities where there is pooling. It

is therefore interesting to know when qR (�) will be locally decreasing, so that the equilibrium

allocation q̂ (�) involves pooling. First, a preliminary result is helpful. Proposition 5 compares the

relaxed allocation to the �rst best allocation, showing that the relaxed allocation is above �rst best

whenever F (�) > F � (�) and is below �rst best whenever F (�) < F � (�):

Proposition 5 Given maintained assumptions:

qR (�)

8>><>>:�� qFB (�)= qFB (�)

< qFB (�)

F (�) > F � (�)

F (�) = F � (�)

F (�) < F � (�)

(�) strict i¤ Cq�qFB (�)

�> 0

Proof. See Section C.1 at the end of this appendix.

Given that F � (�) crosses F (�) once from below, the relationship between the relaxed allocation

and the (strictly increasing) �rst best allocation given in Proposition 5 leads to the conclusion that

qR (�) is strictly increasing near the bottom � and near the top ��. More can be said about pooling

when consumers either have nearly correct beliefs, or are extremely overcon�dent. When consumers�

prior is close to that of the �rm, the relaxed solution is close to �rst best, and like �rst best is

strictly increasing. In this case the equilibrium and relaxed allocations are identical.

When consumers are extremely overcon�dent such that their prior is close to the belief that

� = �� with probability one, the relaxed solution is strictly decreasing at or just above ��. In this

5

Page 6: Web Appendices to Selling to Overcon–dent Consumers

case ironing will be required and an interval of types around �� pool at the same quantity q̂ (��).

The intuition is that when the consumers�prior is exactly the belief that � = �� with probability

one, [F (�)� F � (�)] falls discontinuously below zero at ��. Thus, by Proposition 5, the relaxed

solution must drop discontinuously from weakly above �rst best just below �� to strictly below �rst

best at ��. These results and the notion of "closeness" are made precise in Proposition 6.

Proposition 6 (1) There exists " > 0 such that if jf� (�)� f (�)j < " for all � then qR (�) is

strictly increasing for all �. (2) There exists a �nite constant , such that if f� (��) > and f� (�)

is continuous at �� then qR (�) is strictly decreasing just above ��.

Proof. If neither non-negativity nor satiation constraints bind at �, Lemma 2 and the implicit

function theorem imply that dd�q

R (�) = �q�(qR(�);�)qq(qR(�);�)

. In this case the sign of the cross partial

derivative q� determines whether qR (�) is increasing or decreasing. For any � at which F � is

su¢ ciently close to F in both level and slope, q� > 0, but when f� (�) is su¢ ciently large q� < 0.

See the proof at the end of this appendix for details.

C.1 Pooling Appendix Proofs

C.1.1 Small Lemma 5

Lemma 5 Satiation qS (�) and �rst best qFB (�) quantities are continuously di¤erentiable, strictly

positive, and strictly increasing. Satiation quantity is higher than �rst best quantity, and strictly so

when marginal costs are strictly positive at qFB.

Proof. Satiation and �rst best quantities are strictly positive because by assumption: Vq (0; �) >

Cq (0) � 0. Therefore given maintained assumptions, qS (�) and qFB (�) exist, and are continu-

ous functions characterized by the �rst order conditions Vq�qS (�) ; �

�= 0 and Vq

�qFB (�) ; �

�=

Cq�qFB (�)

�respectively. Zero marginal cost at qFB (�) implies qS (�) = qFB (�). When Cq

�qFB (�)

�>

0, Vqq (q; �) < 0 implies that qS (�) > qFB (�). The implicit function theorem implies dd�q

S = �Vq�Vqq

>

0 and dd�q

FB = � Vq�Vqq�Cqq > 0.

C.1.2 Proof of Proposition 5

Proof. The relaxed allocation maximizes virtual surplus(q; �) within the constraint set�0; qS (�)

�and (q; �) is strictly quasi-concave in q (See proof of Lemma 2). Moreover, qFB 2 (0; qS (�)] andq�qFB (�) ; �

�= Vq�

�qFB (�) ; �

� F (�)�F �(�)f(�) since the �rst best allocation satis�es Vq

�qFB (�) ; �

�=

Cq�qFB (�)

�(Lemma 5). Therefore there are three cases to consider:

1. F (�) = F � (�): In this case virtual surplus and true surplus are equal so qR (�) = qFB (�).

6

Page 7: Web Appendices to Selling to Overcon–dent Consumers

2. F (�) > F � (�): In this case q�qFB (�) ; �

�> 0 and therefore qR (�) � qFB (�). When

Cq�qFB (�)

�= 0, Lemma 5 shows that qS (�) = qFB (�) and therefore the satiation constraint

binds: qR (�) = qFB (�) = qS (�). When Cq�qFB (�)

�> 0, Lemma 5 shows that satiation is

not binding at �rst best, and therefore the comparison is strict: qR (�) > qFB (�).

3. F (�) < F � (�): In this case q�qFB (�) ; �

�< 0 and therefore qR (�) < qFB (�) since, by

Lemma 5, non-negativity is not binding at �rst best (qFB (�) > 0).

C.1.3 Proof of Proposition 6

Proof. If, over the interval (�1; �2), neither non-negativity nor satiation constraints bind and

q (q; �) is continuously di¤erentiable, then in the same interval q�qR; �

�= 0 (Lemma 2) and

the implicit function theorem implies dd�q

R (�) = �q�(qR(�);�)qq(qR(�);�)

. (Given maintained assumptions,

q (q; �) is continuously di¤erentiable at � where f� (�) is continuous.) Therefore, in the same inter-

val qR (�) will be strictly increasing if q��qR (�) ; �

�> 0 and strictly decreasing if q�

�qR (�) ; �

�<

0.

Part (1): As the non-negativity constraint is not binding at � (qR (�) = qFB (�) > 0), the

upper bound qS (�) is strictly increasing, and qR (�) is continuous, qR (�) will be strictly increasing

for all � if the cross partial derivative q� (q; �) is strictly positive for all (q; �) 2�0; qS

��������; ���.

De�ne ' (q; �) and ": (Note that " is well de�ned since�0; qS

��������; ���is compact and F 2 C2

and V (q; �) 2 C3 imply ' (q; �) is continuous.)

' (q; �) � 1

f (�)+(�� � �)f (�)

jVq�� (q; �)jVq� (q; �)

+

�� dd�f (�)

��f (�)

!> 0

" � min(q;�)2[0;qS(��)]�[�;��]

1

' (q; �)

By di¤erentiation:

q� (q; �) = Vq�� (q; �)F (�)� F � (�)

f (�)+ Vq� (q; �)

1 +

f (�)� f� (�)f (�)

�dd�f (�)

f2 (�)[F (�)� F � (�)]

!

As f (�) > 0 and Vq� (q; �) > 0:

q� (q; �) � � jVq�� (q; �)jjF (�)� F � (�)j

f (�)+Vq� (q; �)

1� jf (�)� f

� (�)jf (�)

��� dd�f (�)

��f2 (�)

jF (�)� F � (�)j!

7

Page 8: Web Appendices to Selling to Overcon–dent Consumers

The assumption jf (�)� f� (�)j < " implies that jF (�)� F � (�)j < "(�� � �) and therefore that:

q� (q; �) > Vq� (q; �) (1� "' (q; �))

By the de�nition of ", this implies q� (q; �) > 0 for all (q; �) 2�0; qS

��������; ���.

Part (2): The �rst step is to show that q� (q; ��) < 0. De�ne :

� maxq2[0;qS(��)]

jVq�� (q; ��)jVq� (q; �

�)+ 2f (��) +

�� dd�f (�

�)��

f (��)

!

As f (�) > 0 and Vq� (q; �) > 0:

q� (q; �) � jVq�� (q; �)jjF (�)� F � (�)j

f (�)+ Vq� (q; �)

2� f

� (�)

f (�)+

�� dd�f (�)

��f2 (�)

jF (�)� F � (�)j!

since jF (�)� F � (�)j � 1:

q� (q; �) �Vq� (q; �)

f (�)

jVq�� (q; �)jVq� (q; �)

+ 2f (�) +

�� dd�f (�)

��f (�)

� f� (�)!

By de�nition of and f� (��) > , it follows that for all q 2�0; qS

�����: q� (q; ��) < 0. Given

continuity of f� (�) at ��, q� (q; �) is also continuous at ��. Therefore for some �1 > 0, q� (q; �) < 0

just above �� in the interval � 2 [��; �� + �1).The second step is to show that for some �2 > 0 neither satiation nor non-negativity constraints

are binding in the interval (��; �� + �2). First, satiation is not binding just above �� as qR is below

�rst best here (Proposition 5), which is always below the satiation bound (Lemma 5). Second,

non-negativity is not binding just to the right of �� because qR (��) = qFB (��) > 0 (Proposition 5

and Lemma 5) and qR (�) is continuous (Lemma 2).

Steps one and two imply that qR (�) is strictly decreasing in the interval (��; �� +min f�1; �2g)just above ��. Therefore, qR (�) is either strictly decreasing at �� or has a kink at �� and is

strictly decreasing just above ��. In either case, monotonicity is violated at �� and the equilibrium

allocation will involve pooling at ��.

D Monopoly Multi-Tari¤Menu Extension

In this appendix, I extend the single tari¤ model explored in detail in the main paper to a multi-

tari¤ monopoly model. The model is described in Section 5 of the paper. (Note that I replace

equation (1) by the stricter assumption Vqq� = 0.) The model and solution methods are closely

8

Page 9: Web Appendices to Selling to Overcon–dent Consumers

related to Courty and Li (2000). To work with the new problem, �rst de�ne:

De�nition 2 (1) Let qc(s; �; �0) � minfq(s; �0); qS (�)g be the consumption quantity of a consumerwho chooses tari¤ s, is of type �, and reports type �0. Consumption quantity of a consumer who

honestly reports true type � is qc (s; �) � qc (s; �; �). (2) u(s; �; �0) � V (qc(s; �; �0); �) � P (s; �0)is the utility of a consumer who chooses tari¤ s, is of type �, and reports type �0. (3) u (s; �)

� u (s; �; �) is the utility of a consumer who chooses a tari¤ s, and honestly reports true type �.

(4) U (s; s0) �R ��� u (s

0; �) f (�js) d� is the true expected utility of a consumer who receives signal s,chooses tari¤ s0, and later reports � honestly. U� (s; s0) �

R ��� u (s

0; �) f� (�js) d� is the analogousperceived expected utility. (5) U (s) � U (s; s) is the expected utility of a consumer who honestly

chooses the intended tari¤ s given signal s, and later reports � honestly. U� (s) � U� (s; s) is theanalogous perceived expected utility.

Invoking the revelation principle, the monopolist�s problem may then be written as:

maxq(s;�)�0P (s;�)

E

"Z ��

�(P (s; �)� C (q (s; �))) f (�js) d�

#such that

1. Global IC-2 u (s; �; �) � u(s; �; �0) 8s 2 S; 8�; �0 2 �2. Global IC-1 U� (s; s) � U� (s; s0) 8s; s0 2 S3. Participation U� (s) � 0 8s 2 S

As the signal s does not enter the consumers value function directly, the second period incentive

compatibility constraints may be handled just as they are in the single tari¤ model (See part 1 of

the proof of Proposition 1). In particular, second period local incentive compatibility ( @@�u (s; �) =

V� (qc (s; �) ; �)) and monotonicity ( @@�q

c (s; �) � 0) are necessary and su¢ cient for second period

global incentive compatibility. Moreover, Lemma 1 naturally extends to the multi-tari¤ setting.

Applying the same satiation re�nement q (s; �) � qS (�), the distinction between qc (s; �) and q (s; �)may be dropped.

The next step is to express payments in terms of consumer utility, and substitute in the second

period local incentive constraint, just as was done in the single-tari¤ model (See parts 2 and 3 of

the proof of Proposition 1). This yields analogs of equations (3, 15, 6):

U��s; s0

�= u

�s0; �

�+ E

�V��q�s0; �

�; �� 1� F � (�js)

f (�js)

���� s� (23)

U� (s)� U (s) = E�V� (q (s; �) ; �)

F (�js)� F � (�js)f (�js)

���� s� (24)

9

Page 10: Web Appendices to Selling to Overcon–dent Consumers

P (s; �) = V (q (s; �) ; �)�Z �

�V� (q (s; z) ; z) dz � u (s; �) (25)

In the single tari¤ model, the utility of the lowest type � was determined by the binding partic-

ipation constraint. To pin down the utilities u (s; �) in the multi-tari¤model requires consideration

of both participation and �rst-period incentive constraints given an assumed ordering of the signal

space S. I will consider signal spaces S which are ordered either by FOSD, or a more general reverse

second order stochastic dominance (RSOSD).38 Note that I assume the ordering, either FOSD or

RSOSD, applies to consumer beliefs F � (�js).In the standard single period screening model the participation constraint will bind for the

lowest type, and this guarantees that it is satis�ed for all higher types. The same is true here given

the assumed ordering of S, as stated in Lemma 6.

Lemma 6 If S is ordered by FOSD then the participation constraint binds at the bottom (U� (s) =

0). This coupled with �rst period incentive constraints are su¢ cient for participation to hold for all

higher types s >s. The same is true if S is ordered by RSOSD and V�� � 0.

Proof. It is su¢ cient to show that U� (s; s0) is non-decreasing in s. If this is true, then IC-1

and U� (s) � 0 imply participation is satis�ed: U� (s; s) � U� (s; s) � 0. Hence if U� (s) � 0

were not binding, pro�ts could be increased by raising �xed fees of all tari¤s. Now by de�nition,

U� (s; s0) � E [u (s0; �) js]. By local IC-2 and V� � 0 (which follows from V (0; �) = 0 and Vq� > 0),

it is clear that conditional on tari¤ choice s0, consumers�utility is non-decreasing in realized �:

u��s0; �

�= V�

�q�s0; �

�; ��� 0

Given a FOSD ordering, this is su¢ cient for U� (s; s0) to be non-decreasing in s (Hadar and Russell

1969, Hanoch and Levy 1969).

Taking a second derivative of consumer second period utility shows that conditional on tari¤

choice s0, consumers� utility is convex in � if V�� � 0. This follows from increasing di¤erences

Vq� > 0, monotonicity q� (s0; �) � 0, and local IC-2:

u���s0; �

�= Vq�

�q�s0; �

�; ��| {z }

(+)

q��s0; �

�| {z }(+) by IC-2

+ V���q�s0; �

�; ��| {z }

�0 (assumption)

� 0

38Courty and Li (2000) restrict attention primarily to orderings by �rst order stochastic dominance or meanpreserving spread. However, in their two type model, they do mention that orderings can be constructed from thecombination of the two which essentially allows for the more general reverse second order stochastic dominanceorderings I consider here.

10

Page 11: Web Appendices to Selling to Overcon–dent Consumers

Under RSOSD ordering, this implies that U� (s) is increasing in s. (This result is analogous to the

standard result that ifX second order stochastically dominates Y then E [h (X)] � E [h (Y )] for anyconcave utility h (Rothschild and Stiglitz 1970, Hadar and Russell 1969, Hanoch and Levy 1969).

The proof is similar and hence omitted.)

Given Lemma 6 and the preceding discussion, the monopolist�s problem can be simpli�ed as

described in Lemma 7.

Lemma 7 Given a FOSD ordering of S, or a RSOSD ordering of S and V�� � 0, the monopolist�sproblem reduces to the following constrained maximization over allocations q (s; �) and utilities

u (s; �) for U� (s) = U� (s; s) and U� (s; s0) given by equation (23):

maxq(s;�)2[0;qS(�)]

u(s;�)

E [ (s; q (s; �) ; �)]� E [U� (s)]

such that

1. Monotonicity q (s; �) non-decreasing in �

2. Global IC-1 U� (s; s) � U� (s; s0) 8s; s0 2 S3. Participation U� (s) = 0

(s; q; �) � V (q; �)� C (q) + V� (q; �)F (�js)� F � (�js)

f (�js)

Payments P (s; �) are given as a function of the allocation q (s; �) and utilities u (s; �) by equation

(25). Second period local incentive compatibility ( dd�u (q; �) = V� (q (s; �) ; �)) always holds for the

described payments.

Proof. Firm pro�ts can be re-expressed as shown in (equation 26).

E [� (s; �)] = E [S (s; �)] + E [U� (s)� U (s)]� E [U� (s)] (26)

Substituting equation (24) for �ctional surplus gives the objective function. The participation

constraint follows from Lemma 6. For handling of the satiation constraint and second period

incentive compatibility, refer to discussion in the text and proofs of Lemma 1 and Proposition 1.

To characterize the solution to the monopolist�s problem described by Lemma 7, I now proceed

to analyze a two type (s 2 fL;Hg) model and a continuum of types (s 2 [s; �s]) model separately.

11

Page 12: Web Appendices to Selling to Overcon–dent Consumers

D.1 Two-Tari¤Menu

Assume that there are two possible �rst-period signals s 2 fL;Hg, and that the probability ofsignal H is �. There are two �rst-period incentive constraints. Given Lemma 6, the downward

incentive constraint U� (H;H) � U� (H;L) (IC-H) must be binding. Otherwise the monopolist

could raise the �xed fee P (H; �), and increase pro�ts. Together with equation (23) and the binding

participation constraint U� (L) = 0 (IR-L), this pins down U� (H) as a function of the allocation

q (L; �). Substituting the binding IR-L and IC-H constraints for u (L; �) and u (H; �) into both

the monopolist�s objective function described in Lemma 7 and the remaining upward incentive

constraint U� (L;L) � U� (L;H) (IC-L) delivers a �nal simpli�cation of the monopolist�s problemin Proposition 7.

Proposition 7 De�ne a new virtual surplus � (qL; qH ; �) and function � (qL; qH ; �) by equations

(27) and (28) respectively.

� (qL; qH ; �) ��(H; qH ; �) fH (�) + (1� �) (L; qL; �) fL (�)

��V� (qL; �) (F �L (�)� F �H (�))(27)

� (qL; qH ; �) � [F �L (�)� F �H (�)] [V� (qL; �)� V� (qH ; �)] (28)

Given either a FOSD ordering of S, or a RSOSD ordering and V�� � 0:1. A monopolist�s optimal two-tari¤ menu solves the reduced problem:

maxqL(�);qH (�)2[0;qS(�)]

Z ��

�� (qL (�) ; qH (�) ; �) d�

such that

IC-2 qL (�) and qH (�) non-decreasing in �

IC-L �R ��� � (qL (�) ; qH (�) ; �) d� � 0

2. Payments are given by equations (25), and (29-30):

u (L; �) = �E�V� (qL (�) ; �)

1� F �L (�)fL (�)

����L� (29)

u (H; �) = E

�[V� (qL (�) ; �)� V� (qH (�) ; �)]

1� F �H (�)fH (�)

����H�+ u (L; �) (30)

Proof. Beginning with the result in Lemma 7, the simpli�cation is as follows: Equation (23) and

U� (L) = 0 imply equation (29). Equation (23) and U� (H;H) = U� (H;L) imply equation (30). By

equation (23) and equations (29-30) U� (H) isR ��� V� (qL (�) ; �) [F

�L (�)� F �H (�)] d�. As U� (L) = 0,

this means E [U� (s)] = �R ��� V� (qL (�) ; �) [F

�L (�)� F �H (�)] d�, which leads to the new term in the

12

Page 13: Web Appendices to Selling to Overcon–dent Consumers

revised virtual surplus function �. Further, as U� (L) = 0, the upward incentive constraint (IC-L)

is �U� (L;H) � 0. By equation (23) and equations (29-30) U� (L;H) isR ��� � (qL (�) ; qH (�) ; �) d�

which gives the new expression for the upward incentive constraint.

Given the problem described by Proposition 7, a standard approach would be to solve a relaxed

problem which ignores the upward incentive constraint, and then check that the constraint is

satis�ed. Given a FOSD ordering, a su¢ cient condition for the resulting relaxed allocation to solve

the full problem is for q (s; �) to be non-decreasing in s. (Su¢ ciency follows from IC-H binding and

Vq� > 0.) This is not the most productive approach in this case, because the su¢ cient condition is

likely to fail with high levels of overcon�dence. An alternative approach is to directly incorporate

the upward incentive constraint into the maximization problem using optimal control techniques.

Proposition 8 Given either a FOSD ordering or a RSOSD ordering and V�� � 0: The equilibriumallocations q̂L (�) and q̂H (�) are continuous and piecewise smooth. For �xed � 0, de�ne relaxedallocations (which ignore monotonicity constraints):

qRL (�) = arg maxqL 2[0;qS(�)]

�(L; qL; �)�

+ �

1� �V� (qL; �)F �L (�)� F �H (�)

fL (�)

�(31)

qRH (�) = arg maxqH2[0;qS(�)]

�(H; qH ; �) +

�V� (qH ; �)

F �L (�)� F �H (�)fH (�)

�(32)

The relaxed allocations qRL (�) and qRH (�) are continuous and piecewise smooth functions character-

ized by their respective �rst order conditions except where satiation or non-negativity constraints

bind. Moreover, there exists a non-negative constant � 0 such that:

1. On any interval over which a monotonicity constraint is not binding, the corresponding equi-

librium allocation is equal to the relaxed allocation: q̂s (�) = qRs (�).

2. On any interval [�1; �2] such that the monotonicity constraint is binding inside, but not just

outside the interval for tari¤ s, the equilibrium allocation is constant at some level q̂s (�) = q0

for all � 2 [�1; �2]. Further, the pooling quantity q0 and bounds of the pooling interval [�1; �2]must satisfy qRs (�1) = q

Rs (�2) = q

0 and meet the �rst order condition from the relaxed problem

in expectation over the interval. For instance, for qL (�) this second condition is:Z �2

�1

�q (L; q̂; �)�

+ �

1� �Vq� (q̂; �)F �L (�)� F �H (�)

fL (�)

�fL (�) d� = 0

3. Complementary slackness: = 0 or IC-L binds with equality (R ��� � (q̂L (�) ; q̂H (�) ; �) d� = 0).

13

Page 14: Web Appendices to Selling to Overcon–dent Consumers

Proof. I apply Seierstad and Sydsæter (1987) Chapter 6 Theorem 13, which gives su¢ cient

conditions for a solution. To apply the theorem, I �rst restate the problem described by Proposition

7 in the optimal control framework. This includes translating the upward incentive constraint into a

state variable k (�) =R �� �� (qL (z) ; qH (z) ; z) dz with endpoint constraints k (�) = 0 (automatically

satis�ed) and k(��) � 0. Remaining state variables are qL (�) and qH (�), which have free endpoints.Control variables are cL (�), cH (�), and ck (�), and the control set is R3. Costate variables are

�L (�), �H (�), and �k (�).

maxqL;qH

Z ��

�� (qL (�) ; qH (�) ; �) d�

State Control Costate Control - State Relation

qL (�) cL (�) �L (�) _qL (�) = cL (�)

qH (�) cH (�) �H (�) _qH (�) = cH (�)

k (�) ck (�) �k (�) _k (�) = �� (qL; qH ; �)

End Point Constraints Lagrangian Multipliers

(by de�nition) k (�) = 0

IC-L k(��) � 0 �

Control Constraints Lagrangian Multipliers

Monotonicity cL (�) � 0 �L (�)

cH (�) � 0 �H (�)

State Constraints Lagrangian Multipliers

Non-negativity qL (�) � 0 �L (�), �L, ��L

qH (�) � 0 �H (�), �H , ��H

Satiation�qS (�)� qL (�)

�� 0 �L (�), �L, ��L�

qS (�)� qH (�)�� 0 �H (�), �H , ��H

The Hamiltonian and Lagrangian for this problem are:

H = �(qL; qH ; �) + �L (�) cL (�) + �H (�) cH (�)� �k (�) � (qL; qH ; �)

L = H + �L (�) cL (�) + �H (�) cH (�) + �L (�) qL (�) + �H (�) qH (�)

+�L (�)�qS (�)� qL (�)

�+ �H (�)

�qS (�)� qH (�)

�The following 7 conditions are su¢ cient conditions for a solution:

1. Control and state constraints are quasi-concave in states qL, qH , and k, as well as controls cL,

14

Page 15: Web Appendices to Selling to Overcon–dent Consumers

cH , and ck for each �. Endpoint constraints are concave in state variables. This is satis�ed

because all constraints are linear.

2. The Hamiltonian is concave in states qL, qH , and k, as well as controls cL, cH , and ck for each

�. This is satis�ed because H is constant in k and ck, linear in cL and cH , strictly concave

in qL and qH (equations 33-34), and all other cross partials are zero. This relies on Vqq� = 0

and Sqq (q; �) < 0.

@2H

@q2L= (1� �)Sqq (qL; �) < 0 (33)

@2H

@q2H= �Sqq (qH ; �) < 0 (34)

3. State and costates are continuous and piecewise di¤erentiable in �. Lagrangian multiplier

functions as well as controls are piecewise continuous in �. All constraints are satis�ed.

4.@L̂

@cL=@L̂

@cH=@L̂

@ck= 0

5.

dd��L = �

@L̂@qL; d

d��H = �@L̂@qH

; dd��k = �

@L̂@k

6. Complementary Slackness conditions

(a) For all �:

�L; �H ; �L; �H ; �L; �H � 0

�LcL = �HcH = �LqL = �HqH = �L�qS � qL

�= �H

�qS � qH

�= 0

(b) For ��

��L; ��H ; ��L; ��H � 0

��LqL(��) = ��HqH(

��) = ��L�qS(��)� qL(��)

�= ��H

�qS(��)� qH(��)

�= 0

(c) For �

�L; �H; �L; �H � 0

�LqL (�) = �HqH (�) = �L

�qS (�)� qL (�)

�= �H

�qS (�)� qH (�)

�= 0

(d) For IC-L

� � 0

15

Page 16: Web Appendices to Selling to Overcon–dent Consumers

k (�) = � k(��) = 0

7. Transversality Conditions

�L(��) = ��L � ��L; �H(��) = ��H � ��H ; �k(��) = �

�L (�) = ��L + �L; �H (�) = ��H + �H �k (�) = �

Fix � 0. Let �k (�) = � = � = . As �k (�) is constant, it is continuous and di¤erentiable,and d

d��k = 0. Neither the state k (�) nor the control ck (�) enter the Lagrangian, so@L̂@ck

= �@L̂@k = 0.

For any allocation, k (�) = 0 by de�nition, so k (�) = 0. Putting aside for now the constraint

k(��) � 0 and the complementary slackness condition � k(��) = 0, all other conditions concerning

state k (�) are satis�ed. Moreover, the Hamiltonian and Lagrangian are both additively separable in

qL and qH . Hence the remaining conditions are identical to those for two independent maximization

problems. Namely:

maxqL2[0;qS(�)]

dd�qL�0

�(L; qL; �)�

+ �

1� �V� (qL; �)F �L (�)� F �H (�)

fL (�)

maxqH2[0;qS(�)]

dd�qH�0

�(H; qH ; �) +

�V� (qH ; �)

F �L (�)� F �H (�)fH (�)

For �xed � 0, the solution to these problems exists, satis�es Proposition 8 parts 1-2, and

meets all the relevant conditions in 3-7 above. The proof for this statement is omitted, because the

solution for each subproblem given any �xed � 0, is similar to the single-tari¤ case, and closelyparallels standard screening results. The idea is that for regions (�1; �2) where monotonicity is

not binding for allocation q̂s (�), the q̂s (�) subproblem conditions are the Kuhn-Tucker conditions

for the relaxed solution qRs (�). For regions in which monotonicity is binding, the characterization

closely parallels Fudenberg and Tirole�s (1991) treatment of ironing in the standard screening model.

For the nice properties of the relaxed solution stated in the proposition, refer to the analogous proof

of Proposition 2 part 1.

All that remains to show is that there exists a � 0, such that k(��) � 0 and k(��) = 0 givenqL (�) and qH (�) that solve the respective subproblems for that . If k(��) � 0 for = 0 there

is no problem. If k(��) < 0 for = 0, the result follows from the intermediate value theorem and

two observations: (1) k(��) varies continuously with , and (2) for su¢ ciently large k(��) > 0.

The latter point follows because each subproblem is a maximization of E [� (qL; qH ; �)] + k(��).

As increases, the weight placed on k(��) in the objective increases, and k(��) moves towards its

16

Page 17: Web Appendices to Selling to Overcon–dent Consumers

maximum. Given FOSD, k(��) is maximized (given non-negativity, satiation, and monotonicity) at

fqL (�) ; qH (�)g =�0; qS (�)

, for which qH > qL and therefore, by binding IC-H and Vq� > 0, IC-L

must be strictly satis�ed. The argument is more involved, but the constrained maximum of k(��) is

also strictly positive given RSOSD.

D.2 A Continuum of Tari¤s

For a continuum of �rst-period signals, Courty and Li (2000) characterize a necessary local condition

for �rst period incentive compatibility. Its analog here is based on consumers perceived utility U� (s)

(Lemma 8).

Lemma 8 Tari¤ fq (s; �) ; P (s; �)g satis�es period 1 incentive constraints only if

dU� (s)

ds= �

Z ��

�V� (q (s; �) ; �)

@F � (�js)@s

d�

Proof. By the envelope theorem dU�(s)ds = @U�(s)

@s . Thus dU�(s)ds =

R ��� u (s; �)

@f�(�js)@s d�. Integrating

by parts and substituting the second period local incentive condition @u(s;�)@� = V� (q (s; �) ; �) then

yields the desired result.

Now, applying the fundamental theorem of calculus, substituting the results of Lemmas 6 and 8,

taking expectations, and integrating by parts yields the following expression for expected perceived

utility E [U� (s)]:

E [U� (s)] =

Z �s

s

Z ��

�V� (q (s; �) ; �) (1�G (s))

@

@s(1� F � (�js)) d�ds (35)

Finally, combining equation (35) with Lemma 7, the monopolist�s problem may now be re-

written as given in Proposition 9:

Proposition 9 The monopolist�s problem may be re-expressed as the constrained maximization of

expected virtual surplus � (q; s; �):

maxq(s;�)2[0;qS(�)]

Z �s

s

Z ��

�� (s; q (s; �) ; �) f (�js) g (s) d�ds

such that

1. monotonicity @@�q (s; �) � 0

2. Global IC-1 U� (s; s) � U� (s; s0) 8s; s0 2 S

� (s; q; �) � (s; q; �)� V� (q; �)1�G (s)g (s)

@@s (1� F

� (�js))f (�js)

17

Page 18: Web Appendices to Selling to Overcon–dent Consumers

Proof. Outlined in text.

One can see that virtual surplus now includes the information rent term 1�G(s)g(s)

@@s[1�F �(�js)]f(�js)

which arises in Courty and Li�s (2000) model, as well as the �ctional surplus term F �(�js)�F (�js)f(�js)

which arose in the single tari¤model with overcon�dence. Ignoring the three remaining constraints,

the relaxed solution is characterized by the �rst order condition �q = 0 because virtual surplus is

concave in q (�qq < 0) under the maintained assumption Vqq� = 0.

Given a FOSD ordering, as in Courty and Li�s (2000) model, allocation q (s; �) non-decreasing in

signal s is su¢ cient but not necessary for �rst period global incentive compatibility. Now, however,

general examples for which Courty and Li (2000) were able to show the relaxed solution was non-

decreasing in signal smay violate this su¢ cient condition given high enough levels of overcon�dence.

This is because for high levels of overcon�dence, the relaxed solution involves marginal prices near

the monopoly price for each particular minute as discussed in Web Appendix B. As monopoly price

increases with demand, this implies types with higher signals should face higher marginal prices at

a given quantity, and therefore consume less for a given �. This is unfortunate, because in these

cases it is not known what the optimal tari¤ will look like.39

For speci�c cases in which the allocation of the relaxed solution is non-decreasing in signal s,

marginal price is given by equation (36) if it is well de�ned, where � = � (s; q) is the inverse of

q̂ (s; �) if the latter is invertible.

Pq(s; q) =Max

8<:0; Cq (q) + Vq� (q; �)8<:

1�G(s)g(s)

@@s[1�F �(�js)]f(�js)

+F �(�js)�F (�js)f(�js)

9=;9=; (36)

Marginal pricing for the top tari¤ intended for consumers with signal �s is the same as in the

single tari¤ model. Marginal prices for all lower tari¤s on the menu are distorted upwards by the

information rent term. Whether these tari¤s continue to o¤er initial quantities at zero marginal

price depends on the relative size of the information rent and �ctional surplus. Initial quantities

are more likely to be o¤ered at zero marginal price for all tari¤s on the menu when overcon�dence

is high, and ex ante heterogeneity is low.

E Price Discrimination with Common Priors

The model description and analysis follow directly on from where appendix D.2 left o¤above, simply

by assuming that consumers�priors are correct: F � (�js) = F (�js), and leaving o¤ all superscript

39Global incentive compatibility would need to be checked directly, and if it failed an ironing procedure could notbe used to �nd the optimal tari¤ because monotonicity is not necessary.

18

Page 19: Web Appendices to Selling to Overcon–dent Consumers

�. In this case the virtual surplus function characterized in Proposition 9 reduces to:

� (s; q; �) � V (q; �)� C (q)� V� (q; �)1�G (s)g (s)

@@s (1� F (�js))

f (�js)

It is now straight forward to characterize a solution to the relaxed problem in which the ad-

ditional global incentive constraints are ignored. Virtual surplus is concave in quantity (Equation

38) so the relaxed solution qR (s; �) is characterized by the �rst order condition (Equation 37).

�q (s; q; �) = Vq (q; �)� Cq (q)� Vq� (q; �)(1�G (s))g (s)

@@s (1� F (�js))

f (�js) = 0 (37)

�qq (s; q; �) = Vqq (q; �)| {z }(�)

� Cqq (q)| {z }(+)

� Vqq� (q; �)| {z }=0

(1�G (s))g (s)

@@s (1� F (�js))

f (�js) � 0 (38)

Unfortunately, if the relaxed solution does not satisfy global incentive compatibility, no char-

acterization of the true solution is given. This is because there is no known necessary and su¢ -

cient condition for �rst period incentive compatibility analogous to the monotonicity requirement

q� (s; �) � 0 which can be imposed in an ironing procedure.Nevertheless, there are conditions su¢ cient to verify that the relaxed solution is globally incen-

tive compatible. In particular, given the �rst order local incentive condition dU(s)ds = �

R ��� V� (q (s; �) ; �)

@F (�js)@s d�,

a positive cross partial derivative @2U(s;s0)@s@s0 � 0 is a su¢ cient second order condition to guarantee

global period one incentive compatibility. This gives the following lemma.

Lemma 9 If the relaxed solution qR (s; �) satis�es (1) monotonicity in �: @@�q

R (s; �) � 0 8s; �and (2) @

@sqR (s; �) � @@s [1� F (�js)] � 0 8s; � then q

R (s; �) and the corresponding implicit payment

function solve the original problem.

Proof. Di¤erentiating equation (23), the cross partial is:

@2U (s; s0)

@s@s0=

Z ��

�Vq��q�s0; �

�; �� @@sq�s0; �

� @@s[1� F (�js)] d�

By assumption Vq� is positive, so @@sq (s

0; �) @@s [1� F (�js)] � 0 guarantees that@2U(s;�s)@s@�s � 0. Given

the �rst period local incentive constraint, this serves as a su¢ cient second order condition for

�rst period global incentive compatibility. Finally, @@�q

R (s; �) � 0 guarantees second period globalincentive compatibility.

The conditions given in Lemma 9 are given in terms of the relaxed solution. Using standard

monotone comparative statics results, su¢ cient conditions based on the primitives of the model can

19

Page 20: Web Appendices to Selling to Overcon–dent Consumers

be derived from the cross partials �q� and �qs. Following de�nition 3, these are given in Lemma

10.

De�nition 3

� (s; �) �@@s (1� F (�js))

f (�js)

(Baron and Besanko (1984) refer to this as an "informativeness measure")

Lemma 10 Condition (1) @@�q

R (s; �) � 0 8s; � holds if and only if � (s; �) does not increase in �too quickly:

(1�G (s))g (s)

� @@�� (s; �) � 1 8s; �

If G (s) satis�es the monotone hazard condition ( @@s

�1�G(s)g(s)

�� 0) then condition (2) @@sq

R (s; �) �@@s [1� F (�js)] � 0 8s; � holds if � (s; �) and

@@s� (s; �) have opposite signs:

� (s; �) � @@s� (s; �) � 0

Proof. Follows from inspection of �q� and �qs, Vq� > 0, and standard monotone comparative

statics results.

Given the analysis thus far, it is now possible to consider whether or not menus of three-part

tari¤s of the type o¤ered by cell phone service providers would be predicted by the relaxed solution

of the model.

As absent pooling, the consumer will consume until his marginal valuation is equal to marginal

price (Pq (s; q) = Vq (q; �)), marginal prices under the relaxed solution fall out of the �rst order

condition in equation (37). In particular:

Pq (s; q) = Cq (q) + Vq� (q; �)(1�G (s))g (s)

@@s (1� F (�js))

f (�js)

By inspection, marginal price is above marginal cost whenever @@s (1� F (�js)) is positive, and

marginal price is below marginal cost whenever @@s (1� F (�js)) is negative.

First consider the special case of RSOSD where S is ordered by �rst order stochastic dominance

(FOSD) so that F (�jsH) � F (�jsL) for all � and for all sH � sL. In this case @@s (1� F (�js)) � 0

and therefore marginal prices are everywhere weakly above marginal cost.

The general su¢ cient conditions for global incentive compatibility given in Lemma 10 remain

somewhat opaque. That being said, for a simple parametrization of � it is easy to verify global

incentive compatibility.

20

Page 21: Web Appendices to Selling to Overcon–dent Consumers

Lemma 11 Assume that � is a sum of the signal s and a random variable " with smooth cumulative

density H ("):

� = s+ "

" � H (")

Then F (�js) are ordered by FOSD and if G (s) satis�es the monotone hazard rate condition@@s

�1�G(s)g(s)

�� 0, then the relaxed solution qR (s; �) satis�es global incentive compatibility.

Proof. First, the FOSD ordering follows because F (�js) = H (� � s) is decreasing in s. Second,when � is a function of s and a random variable " � H ("), and � is strictly increasing in ", it canbe shown that @

@�� (s; �) =�s"�"and @

@s� (s; �) = �ss ��s"�"�s. In this parameterization �ss = �s" = 0

which implies @@�� (s; �) =

@@�� (s; �) = 0 and therefore the conditions of Lemma 10 are met.

No absolute conclusions can be drawn without a characterization of the solution when it di¤ers

from the relaxed solution. That being said, the preceding discussion suggests that if S is ordered

by FOSD, the common-prior model cannot explain observed cell phone service pricing.

It is possible, however, to work backwards to see what type distributions might explain observed

pricing. Following the approach taken in this paper, now assume that consumers may freely dispose

of minutes beyond their �nite satiation points. As seen earlier in the paper, this essentially imposes

the constraint that marginal price be non-negative. Further, when marginal costs are assumed to

be zero, this constraint binds whenever @@s (1� F (�js)) is weakly negative.

Observed menus of tari¤s, with increasing levels of included minutes, might then correspond to

the relaxed solution if the RSOSD ordered conditional distributions satisfy equation (39) for some

cuto¤ �� (s) increasing in s. The relaxed solution would then involve zero marginal price up to

qR (s; �� (s)) on each tari¤, and positive marginal prices at higher quantities.

@

@s(1� F (�js))

8<: � 0 � � �� (s)> 0 � > �� (s)

(39)

As the su¢ cient conditions for the relaxed solution to be optimal given in Lemma 10 are

somewhat opaque, it is again useful to turn to a particular parameterization where global incentive

compatibility can be veri�ed.

Lemma 12 Assume that � is a function of the signal s and a mean zero (E ["] = 0) random

variable " with smooth cumulative distribution H (") as given below:

� = z + �s+ s1+ "

21

Page 22: Web Appendices to Selling to Overcon–dent Consumers

�; � 0

Then F (�js) are ordered by RSOSD and @@s (1� F (�js))

8<: � 0 � � �� (s)> 0 � > �� (s)

for some �� (s) which

is constant for � = 0 and strictly increasing for � > 0. Moreover, if G (s) satis�es the monotone

hazard rate condition @@s

�1�G(s)g(s)

�� 0, and s is not too small (s � 1+

g(s) ), then the relaxed solution

qR (s; �) satis�es global incentive compatibility.

Proof. First, RSOSD follows as increasing s involves a mean preserving spread and a FOSD shift.

Second �� (s) is found by setting � (�; s) = �s equal to zero and solving for �. This yields

�� (s) = z + ��

1+

�s and @

@s�� (�) = �

� 1+

�which is strictly positive if � > 0.40

Third @@�� (�; s) =

�s"�e= 1+

s . Thus condition (1) reduces to(1�G(s))g(s) (1 + ) � s. Under the

monotone hazard rate assumption, this inequality is hardest to satisfy for s. Thus the assumption

s � 1+ g(s) is su¢ cient and period two global incentive compatibility is satis�ed.

Fourth, it can be shown that:

@

@s� (s; �) = �1

s� � (s; �)�

s�

Therefore � (s; �) � 0 implies that @@s� (s; �) � 0. So condition (2) of Lemma 10 is satis�ed whenever

� (s; �) � 0. (This implies that qR (s; �) is increasing in s whenever � (s; �) � 0 and corresponds tothe observed fall in overage rates on higher tari¤s.)

Of course this condition will not always be met when � (s; �) < 0. This does not matter, however,

because for � (s; �) < 0 zero marginal costs imply that the satiation (free disposal) constraint is

binding and qR (s; �) = qS (�). This in turn means that @@�q

R (s; �) = 0 and condition (2) of Lemma

9 is satis�ed, which is what counts. Thus �rst period global incentive compatibility is guaranteed.41

As an example that meets the conditions of Lemma 12, take � = 5s + s2" for s � U [2; 3], and" � N (0; 1). In this example �� (s) = 5

2s.

40The critical points �� (s) will be within the support of � given s as long as the support of " is large enough:"� � �

1+ s� .

41Note that for = 0, �rst period incentive compatibility holds without relying on the satiation constraint. In

particular, for = 0, �ss = 0 and so � (s; �) � @@s� (s; �) = �s

��ss � �s"

�"�s�� 0 since �s"; �" > 0. This means that

condition (2) of lemma 10 always holds. The fact that this works out so nicely is related to the fact that �� (s) isconstant in s and therefore all the distributions F (�js) cross once at this common point. (Essentially the case Courtyand Li (2000) focus on). However, this would correspond to a menu of tari¤s which all have the same number ofincluded minutes. To explain observed tari¤s > 0 is required so that �� (s) is increasing.

22