+ All Categories
Home > Documents >  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”†...

 · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”†...

Date post: 30-Jul-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
134
1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information” Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School and Iowa State University) Yu Jin (Shanghai University of Finance and Economics) This Revision: May 5, 2020 1. Introduction This document is the online appendix to “Portfolio Liquidity and Security Design with Private Information” by Peter M. DeMarzo, David M. Frankel, and Yu Jin (2020), henceforth “DFJ”. It is organized as follows. Section 2 formally states the technical results that are previewed in section 3.5.2 of DFJ. Section 3 proves these results as well as the results that are stated without proof in section 3 of DFJ. Section 4 relaxes ASSUMPTION A (monotonicity) in the case of 2 assets and 2 seller types. Section 5 checks the robustness of our results to the alternative asset sale procedure of Maskin and Tirole (1992). 2. Technical Results We first formally state the technical results that we discuss in section 3.5.2 of DFJ. We begin by specifying the continuous and discrete models. The continuous model, denoted “model ∞”, is as follows. The issuer's type ~ t G has full support [0,1] and the final asset value Y has conditional and unconditional support 0, y and conditional distribution function H that satisfies the Hazard Rate Ordering property and Lipschitz-H (defined in DFJ, section 3.5.2). DeMarzo: Stanford, CA 94305-5015; [email protected]. Frankel: 200 Leicester St., Carlton, VIC 3053, Australia; [email protected]. Sadly, our coauthor Yu Jin died in 2018.
Transcript
Page 1:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

1

Online Appendix to “Portfolio Liquidity and

Security Design with Private InformationӠ

Peter M. DeMarzo (Stanford University)

David M. Frankel (Melbourne Business School and Iowa State University)

Yu Jin (Shanghai University of Finance and Economics)

This Revision: May 5, 2020

1. Introduction

This document is the online appendix to “Portfolio Liquidity and Security Design with

Private Information” by Peter M. DeMarzo, David M. Frankel, and Yu Jin (2020),

henceforth “DFJ”. It is organized as follows. Section 2 formally states the technical results

that are previewed in section 3.5.2 of DFJ. Section 3 proves these results as well as the

results that are stated without proof in section 3 of DFJ. Section 4 relaxes ASSUMPTION A

(monotonicity) in the case of 2 assets and 2 seller types. Section 5 checks the robustness

of our results to the alternative asset sale procedure of Maskin and Tirole (1992).

2. Technical Results

We first formally state the technical results that we discuss in section 3.5.2 of DFJ. We

begin by specifying the continuous and discrete models. The continuous model, denoted

“model ∞”, is as follows. The issuer's type ~t G has full support [0,1] and the final asset

value Y has conditional and unconditional support 0, y and conditional distribution

function H that satisfies the Hazard Rate Ordering property and Lipschitz-H (defined in

DFJ, section 3.5.2).

† DeMarzo: Stanford, CA 94305-5015; [email protected]. Frankel: 200 Leicester St., Carlton, VIC

3053, Australia; [email protected]. Sadly, our coauthor Yu Jin died in 2018.

Page 2:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

2

We now define a sequence of discrete models i = 1,2,... that converge to model ∞. Let

( )1i i

N

= and ( )

1i iN

= be any two increasing sequences of positive integers. In model i, let

the gaps between adjacent types t and shocks Z be 1i iN = and i iy N = , respectively.

That is, t lies in 0, , ,1 ,1i i iS = − and Z lies in 0, , , ,i i iy yS = − . By

construction, both gaps i and i converge to zero as i goes to infinity. Let the conditional

distribution of Y in model i be the restriction of the continuous distribution function H to

types in iS and shocks in iS . Similarly, the distribution of t in model i is the restriction

of G to types t in iS .79

Let iE and E denote the expectations operators in models i and , respectively. Let

( ) ( ), min , |i iv D t E D Y t= denote the expected payout of simple debt with face value D

in model i given a type it S . Let ( ) ( ), min , |v D t E D Y t = denote the expected

payout of the same security in model given a type [0,1]t .

Fix equilibria in models i and in which the issuer’s security is simple debt. Let i

tD and

tD be the equilibrium face values of these securities for a given type t. The equilibrium

price of this security in model i, denoted pi(t), is simply the security’s expected payout

( ),i i

tv D t . And the issuer’s expected issuance profit, denoted ( )iu t , is simply the gains

from trade ( ) ( )1 ,i i

tv D t− as competition drives investors’ payoffs to zero. Similarly, in

the continuous model the price ( )p t of the security equals the expected payout

),( tv D t and the issuer’s profit ( )u t

equals the expected gains from trade

( ) )1 ,( tv D t − .

79 That is, in model i, the probability that the type does not exceed some t in

iS is G(t), while the probability,

conditional on a type t, that the cash flow Y does not exceed some y in iS is ( )|H y t .

Page 3:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

3

In the present notation, equation (13) is written as:80

THE CONTINUOUS INITIAL VALUE PROBLEM (CIVP).

0D y= and, for each t, ( )

( )2

0|1

0.1 1

tD

yt

t

H y t dydD

dt H D t

==

− −

Our first result shows that CIVP and its discrete analog have unique solutions:

PROPOSITION 11. Assume Hazard Rate Ordering and Lipschitz-H.

1. There exists a unique function D that satisfies CIVP for v v= . This

function is decreasing and differentiable in the type t, and takes values

in (0, ]y . The associated price and profit functions, p and u , are

decreasing and continuous in the type t.

2. For each discrete model i=1,2,..., there exists a unique, decreasing

function iD with 0

iD y= and satisfying (12) with iv v= for all it S

and .i =

The proof for CIVP runs roughly as follows. The Picard-Lindelöf theorem is the usual tool

for proving the existence and uniqueness of the solution to a differential equation.

Unfortunately, we cannot apply this theorem directly because the differential equation in

CIVP is not Lipschitz continuous in D : it approaches negative infinity as D

approaches y .

We sidestep this difficulty in the following way. We define upper and lower bounds on

D using a modification of CIVP that is Lipschitz continuous with constant k. We then

show that as k grows, these upper and lower bounds approach the same limit, which

satisfies CIVP and thus must be its unique solution D .

80 Integrating by parts, ( ) ( )0

, min , |y

yv D t y D dH y t

== equals ( )

0|

D

yD H y t dy

=− whence

( ) ( )2 20

, | .D

yv D t H y t dy

== − As ( ) ( )| ,H D t D t= by Lipschitz-H, CIVP is equivalent to (13).

Page 4:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

4

Having shown the existence of unique face value functions in both models, we now show

that the face value function in the discrete model converges to that of the continuous model

uniformly in the issuer’s type t. As the equilibrium of the discrete model uniquely satisfies

the intuitive criterion, this supports the use of the continuous equilibrium when convenient.

PROPOSITION 12. Assume HRO and Lipschitz-H. For any 0 there is an i∗ such

that for all models i > i∗ and all types t in [0,1], i

t tD D − .81

The idea of the proof is as follows. For any model i = 1,2,... and constant k > 0, we first

show that any solution iD must lie between fixed upper and lower bounds i

kD and i

kD ,

where these bounds are Lipschitz continuous with constant k. Moreover, these bounds

converge to the aforementioned upper and lower bounds on D as i grows. By the prior

intuition, these bounds on D converge in turn to D as k grows. Thus, by taking i and

k to infinity simultaneously, i

kD and i

kD - and thus iD which lies between them – must

converge to the unique solution D of model .

Our next result gives conditions under which the convergence of the discrete model to the

continuous model is uniform in various parameters. This property can be useful in

applications in which the security design game is preceded by some interaction in which

the issuer also chooses an optimal action, as in Frankel and Jin (2015). In such settings,

the result can help establish that the issuer’s optimal choices in the discrete model are well-

approximated by her optimal choice in the continuous model.

Henceforth, we assume the distribution of types G is continuous with a strictly positive

density g that satisfies the following technical condition:

LIPSCHITZ-G (L-G). There are constants ( )3 4, 0,k k such that for all types t and

t′ in [0,1], ( ) 3 g t k and ( ) ( ) 4g t g t k t t −− .

81 Technically, the function Di is defined only for types t in the discrete set Si. We extend it to all types t in

[0,1] by evaluating it at the highest type in Si that does not exceed t.

Page 5:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

5

We now show that the key functions of the model converge uniformly in the distributions

G and H, the type 0,1 ,t and the cash flow parameter y .82 These key functions are the

face value function iD , price functionip , and the conditional profit function iu . We also

show uniform convergence of the issuer’s unconditional expected issuance profits

( )i i iEu E u t= to its continuous analogue, ( )[ ]Eu E u t = . Finally, let ( )i t denote

the issuer's conditional total profits in model i: the sum of issuance profits ( )iu t and the

conditional expected portfolio return |iE Y t . Let ( )i i iE E t = denote

unconditional expected total profits.83 We show that these converge uniformly to their

continuous counterparts, ( ) ( ) |t u t E Y t = + and ( )=E E t .

PROPOSITION 13. Fix constants k₀, k₁, k₂, k₃, k₄, and y , all in (0,∞). Let H be the

set of conditional distribution functions ( )|H z t that satisfy Hazard Rate Ordering

and Lipschitz-H with constants k0, k1, and k2. Let G be the set of distribution

functions G that satisfy Lipschitz-G with constants k3 and k4. For all ε > 0 there is

an i∗ such that for all models i > i∗, G in G, H in H, y in 0,y , and t in [0,1],

|ωi(t)-ω∞(t)| is less than ε for ω equal to D, p, u, and Π; and |Eωi-Eω∞| is less than ε

for ω equal to u and Π.84

We next show a useful comparative statics property: the issuer chooses a higher face value

when the rate of change function in (13) (which is negative) is smaller in absolute value.

PROPOSITION 10 is essentially a special case of this result (and its proof relies on this result).

PROPOSITION 14. Let the discount factors and lie in ( )0,1 . Assume the

conditional distribution functions H and H satisfy HRO and Lipschitz-H.

suppose that for any given D and t, the rate of change function

82 Uniformity in t does not apply to the expected profit functions Eu and EΠ, as they do not depend on t. 83 This last quantity is especially important in applications: if there is a pregame period, the issuer will act

so as to maximize the sum of EΠi (perhaps multiplied by a discount factor) and any pregame payoff. 84 As in Proposition 14, we extend these functions to all types t in [0,1] by evaluating them at the highest type

in Si that does not exceed t.

Page 6:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

6

( )

( )2

0|1

1 1

D

yH y t dy

H D t

=

− −

in (13) is no larger in absolute value when ( ), H equals

( )ˆ ˆ, H than when it equals ( ), H . Let the face value functions ˆtD and tD solve

equation (13) for ( ), H equal to ( )ˆ ˆ, H and ( ), H , respectively. Then for all t,

ˆt tD D : the issuer does not choose a lower face value under ( )ˆ ˆ, H than under

( ), H .

Finally, we show a homogeneity property that can simplify the analysis of models in which

security design is embedded (e.g., Frankel and Jin (2015)).

COROLLARY 15. The face value functions Di and D , the price functions pi and

p, the profit functions iu , i , u , and

, and the issuer’s expected profits ,iEu

iE , Eu, and E , defined above, are all homogeneous of degree one in the

cash flow parameter y .

3. Proofs

We now give the omitted proofs from section 3 of DFJ, as well as the proofs of the results

of section 2 of this online appendix.

PROOF OF PROPOSITION 6. We first define the Intuitive Criterion in the GSD game.

Consider an equilibrium and any interim asset vector I. If type t sticks to her continuation

strategy, she will get ( ) ( ) ( )( )| , , | ,d

t tu t I U I P I I t p= . If instead she chooses some other

ex-post action ( ),P , for her to lose from this deviation it suffices that her equilibrium

payoff ( )|u t I exceeds her maximum payoff ( ) |I

PE W Y t − from the deviation – or,

equivalently, that

( )| ( ) | .I

Pu t I E W Y t + (16)

Page 7:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

7

The right hand side of (16) is thus the minimum revenue that type t requires to be willing

to deviate to ( ),P .

The Intuitive Criterion for the GSD game states that on seeing the deviation ( ),P

following I, investors must put zero probability on type t if she is never willing to choose

( ),P following I but some other type might be: if condition (16) holds for t but fails for

some other type s. That is:

THE INTUITIVE CRITERION (GSD GAME). An equilibrium of the GSD game with

posterior belief function and outcome u is intuitive if, on seeing any action

( ), ,I P , investors’ posterior probability ( )| , ,t I P is zero for any type t that

satisfies (16) as long as there is some type s for which the inequality is reversed:

for which ( )| ( ) | .I

Pu s I E W Y s +

Let ( ) ( )( )ˆ, , , ,E I P p = be any equilibrium of the GSD game. We will show that the

following equilibrium ( ) ( )( ), , , ,E I P p = is also an equilibrium, and is intuitive if

E is. Let ( )# V denote the length of a vector V.

1. Securitization. In E , the issuer’s interim asset vector ( )1I I= consists of a single

security whose payout equals her cash flow: ( )1I Y Y= . She then chooses an ex-

post action as follows for any given t and interim asset vector I. First, if she deviated

in the prior stage (so that I I ), she chooses the ex-post action ( ) ( )( ),t tP I I that

she would choose in E after choosing I. Else she issues an ex-post action that

consists of her equilibrium revenue cap ( )ˆt I in E together with an ex-post asset

vector consisting of a single security whose payout equals the aggregate payout

( ) ( )( )( )( )ˆ#

1

ˆ ˆtP I j

tjP I I Y

= of her equilibrium ex-post security vector ( )ˆtP I in E.

Page 8:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

8

2. Beliefs. Let ( ), |T P I denote the set of types whose strategies in E instruct

them to choose the ex-post action ( ),P conditional on having chosen some

interim asset vector I. Upon seeing an action ( ), ,I P that is expected in E , the

investors’ posterior probability ( | , , )t I P equals the probability that the issuer’s

type is t conditional on her type being in ( ), |T P I .85 That is, it equals

( ) ( )( )

1

, |s T P Ig t g s

if ( ), |t T P I and zero otherwise. Upon seeing an

action ( ), ,I P that is unexpected in E , ( | , , )t I P equals ( | , , )t I P : its value

in E.

3. Pricing. Prices are given by equation (8) with substituted for : for any action

( ), ,I P , the resulting price vector is ( ) ( )( ), , [ | ] ( | , , ).t

p I P E P I Y t t I P =

Pricing and Beliefs ensure that E satisfies Competitive Pricing and Rational Updating.

As for Payoff Maximization, the payoff of an issuer of type t from taking an action ( ), ,I P

that is unexpected in E equals her payoff from taking this action in E, since, by Pricing

and Beliefs, investors respond with the same price vector. As for expected actions in E ,

we rely on the following claim, whose proof appears below.

CLAIM. Fix an action ( ), ,I P that is expected in E . Let denote the set of

equilibrium actions ( )ˆ, ,I P that are taken in E by any type whose prescribed

action in E is ( ), , .I P 86 Then an issuer of any given type gets the same payoff

in E from taking the action ( ), ,I P as she gets in E from taking any action

( )ˆ, ,I P in .

85 An action ( ), ,I P is expected in an equilibrium if it is selected with positive probability in that

equilibrium.

86 By construction of E , a given type t chooses the same equilibrium revenue cap ( )ˆt I in E as in E.

Thus, every equilibrium action in must involve the same revenue cap that is prescribed in E .

Page 9:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

9

By this Claim, no action in E affords an issuer of a given type more than her equilibrium

payoff in E; and taking her prescribed action in E gives her the same payoff that she gets,

in equilibrium, in E. Thus she will take her prescribed action in E : Payoff Maximization

holds, whence E is an equilibrium as claimed.

PROOF OF CLAIM. Let ( )ˆ, ,I P be any action in . Note first that ( )ˆ, ,I P has

the same aggregate payout function ( ) ( )( )1 1W Y P I Y= as ( ), ,I P and thus the

same aggregate expected payout conditional on the issuer’s type, ( ) | .E W Y t

Hence the Claim holds if ( )ˆ, ,I P raises the same revenue in E as ( ), ,I P raises

in E . We first show that all actions in raise the same revenue in E:

LEMMA. In E, following the equilibrium interim asset choice I , if there are two

ex-post actions that are expected given I and that have the same revenue cap and

aggregate payout function, they must raise the same revenue.

PROOF OF LEMMA. Suppose not: there are types t and t (possibly equal) who,

after choosing I and seeing their types, choose ex-post actions ( ),P and

( ), ,P respectively, that have the same revenue cap and aggregate payout

function, such that ( ),P yields less revenue than ( ),P :

( )( )

( )( )# #

1 1, , , ,

P Pj j

j jp I P p I P

= = . Since the aggregate payout

functions are identical, their conditional expectations are the same conditional on

the issuer’s type being t: ( ) ( )ˆ ˆ

| |I I

P PE W Y t E W Y t = . But then by (7), type t’s

payoff ( )ˆ, , | ,U I P t p from ( ),P exceeds her payoff ( )ˆ, , | ,U I P t p from

( ),P which she is therefore unwilling to choose - a contradiction.

By this Lemma, each action in raises the same issuance revenue R in E. And

if, in E, the issuer chooses action ( )ˆ, ,I P in then, by (6) and (8), investors’

Page 10:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

10

willingness to pay ( ) ( )( )#

1

ˆ ˆ, , , ,d

P j

E jWTP I P p I P

= = for her ex-post assets equals

their estimate ( ) ˆ| ( | , , )tE W Y t t I P of the aggregate payout given what the

action ( )ˆ, ,I P reveals about the issuer’s type. The issuance revenue R in E is

then the minimum of ( )ˆ, ,EWTP I P and the cap . Since, by the above Lemma,

this issuance revenue is the same for any action ( )ˆ, ,I P in , it follows that

either (i) ( )ˆ, ,EWTP I P for every such action or (ii) ( )ˆ, ,EWTP I P for

every such action. Moreover, in case (ii), ( )ˆ, ,EWTP I P must take a common

value ( )EWTP for any action ( )ˆ, ,I P in .

Likewise, if the issuer chooses action ( ), ,I P in E then the amount ( )1 , ,p I P

that investors are willing to pay for her ex-post security is just their estimate

( ) | ( | , , )tE W Y t t I P of this security’s payout conditional on the action

( ), ,I P ; the resulting issuance revenue is the minimum of ( )1 , ,p I P and the cap

.

But for each t, by Beliefs, ( | , , )t I P can be written as ( ) ( )

( ) ( )

, |

, |

1

1

t T P I

s T P Is

g t

g s

. And

the sets ( )ˆ, |T P I of types who, in E, take actions ( )ˆ, ,I P , is a partition of

the set ( ), |T P I of types who take the action ( ), ,I P in E . Thus we can

rewrite ( | , , )t I P as a weighted sum ( )( )

( )ˆ, ,ˆ, ,

ˆ| , ,I P

I P

t I P w

where the

weight ( )

( ) ( )

( ) ( )

ˆ, |

ˆ, ,

, |

1

1

s s T P I

I P

s T P Is

g sw

g s

=

is the probability that the issuer would have

chosen ( )ˆ, ,I P in E conditional on her having chosen ( ), ,I P in E . Hence we

Page 11:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

11

can rewrite ( )1 , ,p I P as ( ) ( ) ( )( )

ˆ, ,ˆ, ,

ˆ| | , ,tI P

I P

w E W Y t t I P

or,

equivalently, as ( ) ( )( )

ˆ, ,ˆ, ,

ˆ, ,EI PI P

w R I P

. Since, the weights ( )ˆ, ,I P

w

sum to one,

the revenue ( ), ,E

R I P raised by the action ( ), ,I P in E equals in case

(i) above (where the revenue raised in E by each ( )ˆ, ,I P in is also ) and the

common value ( )EWTP in case (ii) (when the revenue raised in E by each

( )ˆ, ,I P in is also ( )EWTP ). This proves the Claim.

It remains to show that if E is intuitive, so is E . Suppose E is intuitive. Let ( )|u I and

( )|u I denote the equilibrium payoffs in E and E , respectively, of an issuer of type t who

chooses the interim asset vector I and then follows her continuation strategy in the given

equilibrium. Consider any action ( ), ,I P for which there are types t and s satisfying:

( )| ( ) |I

Pu t I E W Y t + and ( )| ( ) | .I

Pu s I E W Y s + (17)

First assume ( ), ,I P is unexpected in E . As previously noted, the payoff of an issuer

of type t from taking the action ( ), ,I P in E equals her payoff from taking this action in

E. It follows that ( ) ( )| |u v I u v I= for ,v s t= . Hence ( )| ( ) |I

Pu t I E W Y t + and

( )| ( ) |I

Pu s I E W Y s + whence, since E is intuitive, ( )| , ,t I P is zero. But since

( ), ,I P is unexpected in E , Beliefs implies that ( )| , ,t I P equals ( )| , ,t I P

which thus also is zero as required.

Now suppose instead that ( ), ,I P is expected in E : ( ), |T P I is nonempty. By (17),

( )|u t I exceeds ( ) |I

PE W Y t − which is the highest payoff type t can expect from

( ), ,I P . Hence t is not in ( ), |T P I whence, by Beliefs, ( )| , ,t I P is zero. We

conclude that E is intuitive as claimed.

Page 12:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

12

PROOF OF PROPOSITION 7. We first define the notions of equilibrium, intuitive beliefs,

and fair pricing for an EPSD game. Each definition is the natural restriction of the

analogous definition for a GSD game.

EPSD EQUILIBRIUM. A perfect Bayesian equilibrium of the EPSD game is a

security design tS M and price cap t + for each type t, as well as a price

function ˆ ( , )p S and belief function ˆ ( | , )t S , with the following properties:

1. Payoff Maximization: for all t, the issuer’s choice ( , )t tS solves

,ˆmax ( , ) ( ) |S p S E S Y t − subject to S .

2. Competitive Pricing: for any monotone security S and price cap , the

price function ˆ ( , )p S equals ˆmin , [ ( ) | ] ( | , )tE S Y t t S .

3. Rational Updating: the investors’ belief function ˆ ( | , )t S follows Bayes’s

rule when applicable.

FAIR PRICING (EPSD GAME). An equilibrium ( )( )0ˆ ˆ, , ,

T

t t tS p

= of the GSD

game is fairly priced if, for each type t, the price of ex-post security tS equals its

expected value conditional on the issuer’s type: ( )ˆ( , )t t tp S E S Y t = .

The outcome of the EPSD game is the function ( ) ˆ ˆ( , ) ( ) |t t tu t p S E S Y t = − giving the

securitization payoff of each type t.

THE INTUITIVE CRITERION (EPSD GAME). A perfect Bayesian equilibrium

( ) ( ) ( )( )0ˆ ˆ, , , , | ,

T

t t tS p

= of the EPSD game, with outcome ( )u , is intuitive if,

for any security S M and revenue cap + for which ( ) ˆ ( ) |u t E S Y t +

for some type t, ( ) ˆ ( ) |u s E S Y s + implies ˆ( | , ) 0s S = .

Define the following two sets.

Page 13:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

13

1. IAS is the set of intuitive Asset Sale Equilibria ( )( ), ( ), ( , ), ( | , )q p p of the asset

sale game GAS with portfolio (F*, a) in which investor beliefs following any

issuance choice ( ),q p depend only on the quantity choices q and the issuer’s

maximum issuance revenue pq = . Changing notation slightly, we now denote

these beliefs as ( )| ,q pq rather than ( )| ,q p .

2. ISD is the set of intuitive equilibria ( ) ( ) ( )( )0ˆ ˆ, , , , | ,

T

t t tS p

= of the Ex Post

Security Design game GSD.

These two sets are equivalent in the following sense

LEMMA.

1. For any equilibrium e in IAS, there is an equilibrium e in ISD with the same

outcome. If type t sells the quantities q(t) in e, then this type issues the security

( ) ( ) ( )*

tS Y q t F Y= in e .

2. For any equilibrium e in ISD, there is a set of equilibria e in IAS with the same

outcome. If type t issues the security St in e , then this type sells the quantities

tSq in any such equilibrium e.

PROOF OF LEMMA. Consider any equilibrium e in IAS. Its outcome is

( ) ( ) ( ) ( )( ) ( )*,u t q t p q t p t f t = − and investors’ price function is

( ) *, ( ) ( | , )t

p q p p f t t q pq= . We build an equivalent Security Design

Equilibrium e in ISD as follows. Type t issues the monotone security

( ) ( ) ( )*

tS y q t F y= with maximum revenue t equal to ( ) ( )p t q t . This implies, in

particular, that ( ) ( ) ( ) ( ) ( )* *

1 1tS

i i i i i iq q t F y F y y y q t− − = − − = ; collecting terms,

( )tSq q t= . Given any security design choice ( ),S , investors respond with beliefs

ˆ ( | , ) ( | , )SS q = and associated price

Page 14:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

14

( ) ( ) ˆ ˆ, max , | ( | , )t

p S E S Y t t S = ,

whence Competitive Pricing holds in .e A type t issuer’s payoff from the choice

( ),S is ( ) ( ) ( ) ( )* *ˆ , | max , ( | , )S S S

tp S E S Y t q f t t q q f t − = − ,

which is identical to her payoff ( ) ( )*,S Sq p q p f t −

from the issuance choice

( ),Sq p in e where p is any n-vector of asset price caps that gives the same maximum

revenue: .Spq = Moreover, all such asset price cap vectors give the same payoff in

e because – by assumption - investor beliefs and thus issuance revenue depend only on

the quantity vector q and maximum revenue qp . Hence, since ( ) ( )( ),q t p t maximizes

the payoff of type t in e, it follows that ( ),t tS maximizes the payoff of type t in e :

Payoff Maximization holds in e . By restricting to choices made in equilibrium, the

preceding also implies that the outcome ( ) ( ) ( )ˆ ˆ , |t t tu t p S E S Y t = − of e equals

the outcome ( ) ( ) ( ) ( )( ) ( )*,u t q t p q t p t f t = − of e. By construction,

( ) ( )ˆ( | , ) ( | , ) ( | , )tS

t t tS q q t p t = = and the latter is given by Bayes’s Rule

whenever possible. Since the issuer’s behavior is also the same in the two cases,

Rational Updating holds: e is a Security Design Equilibrium. Finally, consider any

security design choice ( ),S in e such that, for some type t,

( ) ( ) ( )*ˆ ( ) | Su t E S Y t u t q f t + = + .

Then following the issuance choice ( ),Sq p in e, investors put zero weight on any type

s for which is less than ( ) ( )*Su s q f s+ which equals ( ) ˆ ( ) |u s E S Y s+ . Hence

ˆ ( | , ) ( | , ) 0Ss S s q = = : e is intuitive.

Consider any equilibrium e in ISD. Its outcome is ( ) ( ) ( )ˆ ˆ , |t t tu t p S E S Y t = − and

the investors’ price function is ( ) ( ) ˆ ˆ, max , | ( | , )t

p S E S Y t t S = . We

build an equivalent Asset Sale Equilibrium e in IAS as follows. (As the price cap vector

will not be unique, this defines a set of equilibria e, as indicated in the statement of this

Page 15:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

15

proposition.) Type t chooses the quantity vector ( ) tSq t q= and any price cap vector

( )p t that satisfies ( ) ( ) tp t q t = . In response to any asset sale choice ( ),q p ,

investors’ posterior beliefs are *ˆ( | , ) ( | , )q p qF pq = .87 The associated price vector

is *( , ) ( ) ( | , )

tp q p p f t t q p= , which implies Competitive Pricing in e. The

price functions in e and e are thus related by ( )*ˆ( , ) ,p q p q p qF pq= . Hence, a type t

issuer’s payoff ( ) ( )*,q p q p f t − in e from the issuance choice ( ),q p equals her

payoff ( ) ( )* *ˆ , |p qF pq E qF Y t − in e from the security design ( )*,qF pq . Thus,

since ( ),t tS maximizes the payoff of type t in e , it follows that ( ),tSq p maximizes

the payoff of type t in e, where p is any price cap vector that satisfies tS

tpq = : Payoff

Maximization holds in e. By restricting to choices made in equilibrium, the preceding

also implies that the outcome ( ) ( ) ( ) ( )( ) ( )*,u t q t p q t p t f t = − of e equals the

outcome ( ) ( ) ( )ˆ ˆ , |t t tu t p S E S Y t = − of e . By construction,

( ) ( ) ( ) ( ) ( )*ˆ ˆ( | , ) ( | , ) ( | , )t tq t p t q t F p t q t S = =

and the last is given by Bayes’s Rule whenever possible. Since the issuer’s behavior is

also the same in the two settings, Rational Updating holds: e is an Asset Sale

Equilibrium. Finally, consider any issuance choice ( ),q p in e such that, for some type

t, ( ) * ˆ( ) ( ) ( ) |pq u t qf t u t E S Y t + = + where S denotes the security qF*. Then,

following the corresponding issuance choice ( ),S pq in e , investors put zero weight

on any type s for which ( ) ( ) ( )*ˆ ( ) |pq u s E S Y s u s qf s + = + . Hence,

*ˆ( | , ) ( | , ) 0s q p s qF pq = = : e is intuitive.

Now let *

ASe denote the intuitive equilibrium e* of the Asset Sale game, specialized to the

case in which the endowment ( ),a f equals the endowment ( )* *,a f of GAS. We claim

87 The security qF*(Y) in GSD promises the same realized payout to investors as the quantity vector q in GAS.

Page 16:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

16

that *

ASe lies in IAS. Why? In *

ASe , each type t can choose any price cap vector that is not

less than f*(t) and investors ignore these caps. Thus, beliefs on the equilibrium path do not

depend on the price cap vector. Following any out-of-equilibrium choice ( ),q p , investors’

beliefs are given by (3) and (4), which are also independent of the price cap vector. So

investor beliefs in *

ASe do not depend on the price cap vector at all. Thus, as *

ASe is intuitive

by PROPOSITION 1, it must lie in IAS.

By the above Lemma, then, there exists an equilibrium *

SDe in ISD with the same outcome

as *

ASe , in which each type t issues the security ( ) ( ) ( )* * *

tS Y q t F Y= where ( )*q t is

chosen by type t in *

ASe and solves RLP. Finally, FOSD together with the fact that F* is

monotone implies that ASSUMPTION A holds. Hence, by PROPOSITION 1, the quantity

choice function in any equilibrium in IAS must be a solution ( )*q to RLP. Thus, by the

above Lemma, in any equilibrium in ISD, the issuer’s optimal security *( )tS Y must equal

* *( ) ( )q t F Y where q(t) solves RLP.

PROOF OF PROPOSITION 8. From PROPOSITION 7, ( ) ( ) ( )* * *

tS Y q t F Y= . By PROPOSITION

5, IIS holds so that from PROPOSITION 3, ( )* 1iq t = for i h(t) and ( )* 0iq t = for i h(t).

Therefore ( ) ( ) ( )* * *

t i i iS y q t F y y= = for i h(t) and

( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )( )* * * *

1 1t i i h t h t h t h tS y q t F y y q t y y

− −= = + −

for i h(t). Thus, ( ) ( )** min ,t tS Y D Y= where ( ) ( ) ( ) ( ) ( )( )*

1

*

1h t h t h t h ttD y q t y y− −

= + − . Finally,

since q and h are decreasing in t by PROPOSITION 3, *

tD is decreasing in t. That *

tD

satisfies (11) follows from PROPOSITION 4. Finally, as the face value function is strictly

decreasing, the equilibrium is fully revealing; hence, *( ) (1 ) ( , )tu t v D t= − . Off-

equilibrium, suppose debt * *

1( , )t tD D D+ is issued. The price p that would make type t

willing to deviate satisfies * *( ) ( , ) (1 ) ( , ) ( , )tp u t v D t v D t v D t = + = − + . For any type

s t , the incentive constraint for s implies * * *( , ) ( , ) ( )t tv D t v D s u s− . Combining these,

Page 17:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

17

we have ( )* * *

0

( , ) ( , ) ( , ) ( , ) ( ) ( , )t tp v D t v D t v D s v D s u s v D s

+ − − − +

, where the term

in square brackets is positive: the value of junior debt with face value *

tD D− is higher for

type t than for type s t . Therefore, types s t would not find it profitable to deviate

given price p . The beliefs specified in (3) must therefore concentrate on t , and so the

market price following a deviation to D is ( , )v D t .

Page 18:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

PROOF OF PROPOSITION 9. Let the range of D∞ be (D,y]. Define U(t, t,D

)to be the

payoff E[min(D,Y ) |t

]− δE [min(D,Y ) |t] of an issuer with type t when she issues debt

with face value D and the investors believe her type is t. Then (96)

UD

Ut=

Pr(Y > D|t

)−δ Pr(Y > D|t)

∂ t E[min(D,Y ) |t

]is nonincreasing in t by First-Order Stochastic Dominance (FOSD). Since by assumption

Dt is decreasing in t and since U2(t, t,D

)≥ 0 by FOSD, part 1 of Theorem 6 in Mailath

and von Thadden (2013) implies that an issuer of type t will not imitate any other type:

she will not sell debt with face value D∞

t for any t 6= t. It remains to consider whether she

will sell debt with a face value not in the range of D∞, as well as other types of monotone

securities. We may assume that investors respond to any such deviation with the the most

pessimistic beliefs: that her type is zero.

First consider a deviation of a type t to debt with a face value D that is not in the range

[D,y] of D∞. Assume this deviation makes type t strictly better off than sticking to the

equilibrium. First, suppose D > y. This security has the same payout for any Y as debt

with face value y, and both lead investors to believe that t = 0. By the preceding result,

no type strictly prefers such a deviation. Now consider D < D. Investor beliefs cannot

be more optimistic than those that result from debt with face value D, since these are the

beliefs that the type is one. Moreover, there are gains from trade, so a higher face value

is more profitable for the issuer (holding investor beliefs constant). Hence, any D < D is

worse for the issuer than D = D.

Finally, we consider deviations of type t to a general monotone security S. The issuer’s

securitization profit from S equals the price E[S (Y ) |0

]assigned by investors less the

discounted expected security payout δE[S (Y ) |t

]. This profit can be written as

∫ y

Y=0S (Y )d [H (Y |0)−δH (Y |t)]

where H denotes the conditional distribution of Y given t. The security S is monotone if

and only if S (0) = 0 and, for all Y , the control variable c(Y ) = S′ (Y ) lies in [0,1]. For all

18

Page 19:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Y ∈ [0,y], define the Hamiltonian

H = λ (Y )c(Y )+ S (Y )[H ′ (Y |0)−δH ′ (Y |t)

]+µ0 (Y )c(Y )+µ1 (Y ) [1− c(Y )]

where µ0,µ1 ≥ 0 are Lagrange multipliers that capture the constraints on c and λ (Y ) is

a costate variable. By Pontryagin’s maximization principle (Pontryagin et al 1962), the

optimal control c must maximize the Hamiltonian H while the multipliers µ0,µ1 ≥ 0

must minimize it. As H is linear in c, this implies the first order condition

0 =∂

∂c(Y )H = λ (Y )+µ0 (Y )−µ1 (Y ) (18)

as well as the Kuhn-Tucker (complementary slackness) conditions:

µ0 (Y )c(Y ) = 0 and µ1 (Y ) [1− c(Y )] = 0. (19)

Additionally, for all Y , the costate equation

λ′(Y ) =−HS =−H ′ (Y |0)+δH ′ (Y |t) (20)

must be satisfied. Finally, as the final state S(y) is not fixed, the terminal costate must be

zero:

λ (y) = 0. (21)

Solving (20) subject to (21), we obtain λ (Y ) = [1−H (Y |0)][1−δ

1−H(Y |t)1−H(Y |0)

]. By HRO,

1−H(Y |t)1−H(Y |0) is increasing in Y since t > 0. Hence, by (18), (19), and the nonnegativity of

µ0 and µ1, there exists an D ∈ ℜ such that c(Y ) = 1 for all Y < D and c(Y ) = 0 for all

Y > D: the optimal security is debt with face value D and thus, as shown above, D = D∞t .

Q.E.D.PROPOSITION 9

PROOF OF PROPOSITION 10: First, 1−H (y|t) can be written as 1−H (y|1)+∫ 1

s=t H2 (y|s)ds,

which is smaller for H = H than for H = H. Hence, the rate of change function in (1) is no

larger in absolute value when H equals H than when it equals H. The result then follows

from PROPOSITION 14. Q.E.D.PROPOSITION 10

19

Page 20:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

PROOF OF PROPOSITIONS 11-14, AND COROLLARY 15: In this proof, we will write

D∞ (t) and Di (t) in place of D∞t and Di

t . We will also treat y as a parameter and the relative

cash flow Z = Y/y as an exogenous random variable with support [0,1] and conditional

distribution function H defined by H (z|t) =H (zy|t) for z in [0,1]. Lipschitz-H then entails

the following property.1

LIPSCHITZ-H. There are constants k0, k1, k2 ∈ (0,∞) such that for all z, t, t ′, t ′′ ∈ [0,1],∂ H(z|t)

∂ z ∈(

k0, k1

),−∂ H(z|t)

∂ t ∈[k0z(1− z) , k1

),∣∣∣∂ H(z|t ′)

∂ z − ∂ H(z|t ′′)∂ z

∣∣∣< k2 |t ′− t ′′| , and∣∣∣∣∣ ∂ H (z|t)∂ t

∣∣∣∣∣t=t ′− ∂ H (z|t)

∂ t

∣∣∣∣∣t=t ′′

∣∣∣∣∣< k2∣∣t ′− t ′′

∣∣ .Henceforth we rename H to H and each kn to kn (for n = 0,1,2), and refer to Lipschitz-H

as Lipschitz-H or L-H.

Recall that E i and E∞ denote the expectation operators in models i and ∞. For any

D ∈ℜ, let

vHyi (D, t) = E i [minD,yZ|t] =1/∆′i

∑c=1

min

D,yc∆′i[

H(c∆′i|t)−H

((c−1)∆

′i|t)]

(22)

1Lipschitz-H implies that there are constants k0,k1,k2 ∈ (0,∞) such that for all z, t, t ′, t ′′ ∈ [0,1],

∂ H (z|t)∂ z

=∂H (zy|t)

∂ z=

∂H (zy|t)∂ (zy)

d (zy)dz∈ (k0y,k1y) ,

−∂ H (z|t)∂ t

=−∂H (zy|t)∂ t

∈ [k0y(y− y) ,k1) =[y2k0z(1− z) ,k1

),

∣∣∣∣∣∂ H (z|t ′)∂ z

− ∂ H (z|t ′′)∂ z

∣∣∣∣∣=∣∣∣∣∂H (zy|t ′)

∂ (zy)− ∂H (zy|t ′′)

∂ (zy)

∣∣∣∣ d (zy)dz

=

∣∣∣∣∂H (y|t ′)∂y

− ∂H (y|t ′′)∂y

∣∣∣∣y < k2y∣∣t ′− t ′′

∣∣ ,and ∣∣∣∣∣ ∂ H (z|t)

∂ t

∣∣∣∣∣t=t ′− ∂ H (z|t)

∂ t

∣∣∣∣∣t=t ′′

∣∣∣∣∣=∣∣∣∣ ∂H (zy|t)

∂ t

∣∣∣∣t=t ′− ∂H (zy|t)

∂ t

∣∣∣∣t=t ′′

∣∣∣∣=

∣∣∣∣ ∂H (y|t)∂ t

∣∣∣∣t=t ′− ∂H (y|t)

∂ t

∣∣∣∣t=t ′′

∣∣∣∣< k2∣∣t ′− t ′′

∣∣ .This implies Lipschitz-H with constants k0 = k0 min

y,y2

, k1 = k1 maxy,1, and k2 = k2 maxy,1.

20

Page 21:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

and

vHy (D, t) = E∞ [minD,yZ|t] =∫ 1

z=0minD,yzdH (z|t) (23)

denote the expected payout of debt with face value D in models i and ∞, respectively,

conditional on the issuer’s type t. Integrating by parts yields

vHy (D, t) = D− y∫ D/y

z=0H (z|t)dz. (24)

Given the above notation, the incentive compatibility condition in model i can be stated

as the following initial value problem (henceforth “IVP”):

DISCRETE IVP WITH PARAMETERS H,δ ,y (DPHδy). The condition

vHyi(

DHδyi (t +∆i) , t +∆i

)−δvHyi

(DHδyi (t +∆i) , t

)= (1−δ )vHyi

(DHδyi (t) , t

)(25)

with DHδyi : Si→ℜ, together with the initial value DHδyi (0) = y.

Define the function

f Hδy (D, t) =− 11−δ

vHy2 (D, t)

vHy1 (D, t)

=y

1−δ

∫ Dy

z=0∂H(z|t)

∂ t dz

1−H(

Dy |t) ≤ 0, (26)

where the second equality follows from (24). A restatement of CP in model ∞ is as follows.

CONTINUOUS IVP WITH PARAMETERS H,δ ,y (CPHδy). The differential

equationdDHδy

dt= f Hδy

(DHδy, t

), (27)

with DHδy : [0,1]→ℜ, together with the initial value DHδy (0) = y.

In model i, let pHδyi (t) denote vHyi(

DHδyi (t) , t)

: the price and equilibrium expected

payout of a standard debt contract, conditional on the type t. Let uHδyi (t) denote (1−δ ) pHδyi (t):

the issuer’s profit and expected gains from trade from such a contract conditional on the

type t. Let EuGHδyi = E i[uHδyi (t)

]denote the unconditional expected gains from trade

21

Page 22:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

and issuer’s profit. Analogously, in the continuous model let pHδy (t) denote vHy(

DHδy (t) , t)

,

let uHδy (t) denote (1−δ ) pHδy (t), and let EuGHδy denote E∞

[uHδy (t)

]. Let ΠGHδyi (t) =

uGHδyi (t)+E i [yZ|t] denote the issuer’s conditional (on t) expected total profits in model

i and let EΠGHδyi = E i[ΠGHδyi (t)

]denote her unconditional expected total profits. In

the continuous model, define the analogous quantities ΠGHδy (t) = uGHδy (t) +E∞ [yZ|t]

and EΠGHδy = E∞

[ΠGHδy (t)

]. As noted, we extend the functions in model i to any type

t ∈ [0,1] by evaluating them them at the highest type in Si that does not exceed t.

With this notation, Propositions 11-14, together with Corollary 15, are combined as

follows.

Theorem 1. Fix constants k0, k1, k2, k3, k4, and y, all in (0,∞). Let G be the set of

distribution functions G that satisfy Lipschitz-G with constants k3 and k4. Let H be the

set of conditional distribution functions H that satisfy Lipschitz-H with constants k0, k1,

and k2, and Hazard Rate Ordering. For any distribution function G in G , conditional

distribution function H in H , parameter y in (0,y], and discount factor δ in (0,1):

1. There exists a unique function DHδy that satisfies CPHδy. This function is decreasing

and differentiable, and takes values in (0,y]. The associated price and profit functions,

pHδy and uHδy, are continuous and decreasing in the type t as well.

2. For each discrete model i = 1,2, ..., there exists a unique, decreasing function DHδyi

that satisfies DPHδyi.

3. The sequences of face value functions, price functions, conditional and unconditional

expected securitization profit functions, and conditional and unconditional expected

total profit functions in the discrete models converge to their continuous counterparts

as i grows, uniformly in the distributions G and H, the parameter y, and (except in

the case of the unconditional expected profit functions which do not depend on the

type) the type t ∈ [0,1]. More precisely:

• For all ε > 0 there is an i∗ such that for all models i > i∗, G in G , H in H , y

22

Page 23:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

in (0,y], and t ∈ [0,1],∣∣∣ωHδyi (t)−ωHδy (t)

∣∣∣ < ε for each ω = D, p,u,Π, and∣∣∣EωHδyi−EωHδy∣∣∣< ε for each ω = u,Π.

4. All of the functions defined above are homogeneous of degree one in the parameter

y: ωHδy = yωHδ1 and ωHδyi = yωHδ1i for each ω = D, p,u,Eu,Π,EΠ.

5. Let H, H ∈H and δ , δ ∈ (0,1) satisfy f Hδy ≤ f Hδy ≤ 0. Then DHδy (t)≤DHδy (t)

for all t ∈ [0,1].

We now prove Theorem 1. Without loss of generality, we restrict to face values D that

do not exceed the maximum final asset value y.2 First, we define an integration by parts

formula for the function vHyi defined in (22).

Claim 1. For any face value D ∈ [0,y] and type t,

vHyi (D, t) = D− y∫ D

y

z=0H(

∆′i

⌊z∆′i

⌋|t)

dz. (28)

PROOF OF CLAIM 1. One can easily verify (using H (1|t) = 1 and H (0|t) = 0) that for

D ∈ [0,y], vHyi (D, t) = minD,y−∑

1∆′i−1

c=1 H (c∆′i|t) [minD,y(c+1)∆′i−minD,yc∆′i]

by (22). As c is an integer, c≤ x iff c≤ bxc. Thus,

min

D,y(c+1)∆′i−min

D,yc∆

′i=

y∆′i if c≤

⌊D

y∆′i

⌋−1

D− yc∆′i if c =⌊

Dy∆′i

⌋0 if c >

⌊D

y∆′i

⌋ ,

whence, as D≤ y,

vHyi (D, t) = D− y∆′i

∑bD/(y∆′i)c−1c=1 H (c∆′i|t)

+H(⌊

Dy∆′i

⌋∆′i|t)(

Dy∆′i−⌊

Dy∆′i

⌋)1(⌊

Dy∆′i

⌋≤ 1

∆′i−1) . (29)

2Choosing a higher face value is equivalent to choosing y since the underlying assets cannot be worth

more than y.

23

Page 24:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

If z ∈ [c∆′i,(c+1)∆′i), then bz/∆′ic= c. So

bD/(y∆′i)c−1

∑c=1

H(c∆′i|t)=bD/(y∆′i)c−1

∑c=1

1∆′i

∫ (c+1)∆′i

z=c∆′i

H(

∆′i

⌊z∆′i

⌋|t)

dz

=1∆′i

∫∆′i

⌊D

y∆′i

⌋z=∆′i

H(

∆′i

⌊z∆′i

⌋|t)

1(⌊

z∆′i

⌋<

1∆′i

)dz (30)

since 1(⌊

z∆′i

⌋< 1

∆′i

)equals one for all z < ∆′i

⌊D

y∆′i

⌋. Moreover,

H(⌊

Dy∆′i

⌋∆′i|t)(

Dy∆′i−⌊

Dy∆′i

⌋)1(⌊

Dy∆′i

⌋≤ 1

∆′i−1)

=1∆′i

∫ Dy

z=∆′i

⌊D

y∆′i

⌋H(⌊

Dy∆′i

⌋∆′i|t)

1(⌊

Dy∆′i

⌋≤ 1

∆′i−1)

dz (31)

=1∆′i

∫ Dy

z=∆′i

⌊D

y∆′i

⌋H(⌊

z∆′i

⌋∆′i|t)

1(⌊

z∆′i

⌋<

1∆′i

)dz (32)

where the last equality holds for two reasons. First,⌊

z∆′i

⌋=⌊

Dy∆′i

⌋in the interval of

integration in line (31). Second, since⌊

z∆′i

⌋and 1

∆′iare both integers,

⌊z∆′i

⌋< 1

∆′iif and

only if⌊

z∆′i

⌋≤ 1

∆′i−1. Combining (29), (30), and (32), we obtain

vHyi (D, t) = D− y∫ D

y

z=∆′i

H(

∆′i

⌊z∆′i

⌋|t)

1(⌊

z∆′i

⌋<

1∆′i

)dz.

As 1(⌊

z∆′i

⌋< 1

∆′i

)equals one except possibly at the upper endpoint of the integral, it can

be omitted. Finally, the lower limit of integration can be reduced to zero since H (0|t) = 0.

Q.E.D.Claim 1

We next show that the first derivatives with respect to D and t of the expected payout

vHyi (D, t) in model i are bounded above and below (parts 1 and 2) and converge uniformly

to the respective first derivatives of the expected payout vHy (D, t) in the continuous case

(parts 3 and 4):

Claim 2. Assume L-H. Define

ΩHyi (t ′, t ′′,D′,D′′)= vHyi (D′, t ′)− vHyi (D′′, t ′′)− [vHy (D′, t ′)− vHy (D′′, t ′′)] .

24

Page 25:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

1. For all D ∈ (y∆′i,y], and all t ′, t ′′ in Si such that t ′ > t ′′,

max

0,k0

(D′)2 (3y−2D′)6y2

<

vHyi (D, t ′)− vHyi (D, t ′′)t ′− t ′′

< yk1 min

Dy,1−∆

′i

,

where D′ = D−2y∆′i.

2. For all D′,D′′ in [0,y] such that D′ > D′′ and for all t ∈ Si,

k1

(1− D′′

y+∆i

)>

vHyi (D′, t)− vHyi (D′′, t)D′−D′′

> k0

(1−∆

′i

⌈D′

y∆′i

⌉+∆

′i

)≥ k0

(1− D′

y

).

3. For all ε > 0 there exists an i∗ such that if i> i∗, then for all G∈G , H ∈H , y∈ (0,y],

D in [0,y] and t ′, t ′′ in Si such that t ′ > t ′′,∣∣ΩHyi (t ′, t ′′,D,D)

∣∣< ε (t ′− t ′′) .

4. For all ε > 0 there exists an i∗ such that if i> i∗, then for all G∈G , H ∈H , y∈ (0,y],

t ∈ Si, and D′,D′′ in [0,y] such that D′ > D′′,∣∣ΩHyi (t, t,D′,D′′)

∣∣< ε (D′−D′′) .

PROOF OF CLAIM 2. Part 1. By (28),

vHyi (D, t ′)− vHyi (D, t ′′

)= y

∫ Dy

z=0

[H(

∆′i

⌊z∆′i

⌋|t ′′)−H

(∆′i

⌊z∆′i

⌋|t ′)]

dz

which by L-H is less than yk1 (t ′− t ′′)(

Dy −∆′i

)and at least

yk0(t ′− t ′′

)(∆′i)2∫ D

y

z=0

⌊z∆′i

⌋(1∆′i−⌊

z∆′i

⌋)dz, (33)

which is zero if D ≤ y∆′i and positive otherwise. Let c =⌊

Dy∆′i

⌋where D′ = D− 2y∆′i.

Recall N′i = 1/∆′i. Hence, for D > y∆′i, the integral in (33) is at least∫ c∆′i

z=0

⌊z∆′i

⌋(N′i −

⌊z∆′i

⌋)dz = ∆

′i

c−1

∑n=1

n(N′i −n

)= ∆

′ic(c−1)(3N′i −2c+1)

6

≥ ∆′i

(D′y∆′i

+1)

D′y∆′i

(3N′i −2 D′

y∆′i−1)

6as c≥ D′

y∆′i+1

>(D′)2 (3y−2D′)

6y3(∆′i)2 as D′ ≤ y

(1−2∆

′i).

25

Page 26:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

This proves the result.

Part 2. By L-H and (22),

vHyi (D′, t)− vHyi (D′′, t)=

1/∆′i

∑c=1

[min

D′,yc∆

′i−min

D′′,yc∆

′i][

H(c∆′i|t)−H

((c−1)∆

′i|t)]

but

min

D′,yc∆′i−min

D′′,yc∆

′i=

0 if c≤ D′′

y∆′i

yc∆′i−D′′ if D′′y∆′i

< c < D′y∆′i

D′−D′′ if c≥ D′y∆′i

=

0 if c≤

⌊D′′y∆′i

⌋yc∆′i−D′′ if

⌊D′′y∆′i

⌋< c <

⌈D′y∆′i

⌉D′−D′′ if c≥

⌈D′y∆′i

⌉so

vHyi (D′, t)− vHyi (D′′, t)≥ 1/∆′i

c=⌈

D′y∆′i

⌉(D′−D′′)[

H(c∆′i|t)−H

((c−1)∆

′i|t)]

>(D′−D′′

)k0∆′i

(1∆′i−⌈

D′

y∆′i

⌉+1)

≥(D′−D′′

)k0(1−D′/y

).

Moreover, vHyi (D′, t)− vHyi (D′′, t) is at most

1/∆′i

c=⌈

D′′y∆′i

⌉(D′−D′′)[

H(c∆′i|t)−H

((c−1)∆

′i|t)]

<(D′−D′′

)k1∆′i

(1∆′i−⌈

D′′

y∆′i

⌉+1)≤(D′−D′′

)k1(1−D′′/y+∆i

).

26

Page 27:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Part 3. Let i be large enough that yk2∆′i < ε . Then

∣∣∆(t ′, t ′′,D,D)∣∣= ∣∣∣∣∣1/∆′i

∑c=1

∫ c∆′i

z=(c−1)∆′i

[min

(D,yc∆

′i)−min(D,yz)

]d[H(z|t ′)−H

(z|t ′′)]∣∣∣∣∣

≤1/∆′i

∑c=1

∫ c∆′i

z=(c−1)∆′i

∣∣min(D,yc∆

′i)−min(D,yz)

∣∣d [H (z|t ′)−H(z|t ′′)]

≤1/∆′i

∑c=1

yk2(∆′i)2 ∣∣t ′− t ′′

∣∣= yk2∆′i∣∣t ′− t ′′

∣∣< ε∣∣t ′− t ′′

∣∣since |min(D,yc∆′i)−min(D,yz)| < y∆′i and by L-H,

∣∣∣∫ c∆′iz=(c−1)∆′i

d [H (z|t ′)−H (z|t ′′)]∣∣∣ ≤

k2∆′i |t ′− t ′′|.

Part 4. As y > 0,∣∣ΩHyi (t, t,D′,D′′)

∣∣ = ∣∣∣∑1/∆′ic=1

∫ c∆′iz=(c−1)∆′i

η (yc∆′i,yz,D′′,D′)dH (z|t)∣∣∣

where η (ζ ′,ζ ′′,ζ0,ζ1) = maxζ0,minζ1,ζ′−maxζ0,minζ1,ζ

′′.

Remark 1. η (ζ ′,ζ ′′,ζ0,ζ1) lies in [0,ζ ′−ζ ′′] if ζ e′ ≥ ζ e′′ and is zero if either

max

ζ′,ζ ′′

≤ ζ e0

or minζ ′,ζ ′′ ≥ ζ e1.

Since z lies in [(c−1)∆′i,c∆′i], the integrand η (yc∆′i,yz,D′′,D′) is zero if either c≤ D′′y∆′i

or c≥ D′y∆′i

+1. Hence,

1/∆′i

∑c=1

∫ c∆′i

z=(c−1)∆′iη(yc∆′i,yz,D′′,D′

)dH (z|t)

=

min

1,⌊

D′y∆′i

+1⌋

c=⌈

D′′y∆′i

⌉∫ c∆′i

z=(c−1)∆′iη(yc∆′i,yz,D′′,D′

)dH (z|t) . (34)

Since c∆′i ≥ z in each integral, the integrands η (yc∆′i,yz,D′′,D′) are all nonnegative, so we

may dispense with the absolute value signs. The first summand, which corresponds to

27

Page 28:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

c = dD′′/y∆′ie, is

∫ ⌈D′′y∆′i

⌉∆′i

z=(⌈

D′′y∆′i

⌉−1)

∆′i

η

(y⌈

D′′

y∆′i

⌉∆′i,yz,D′′,D′

)dH (z|t)

=

η

(y⌈

D′′y∆′i

⌉∆′i,D

′′,D′′,D′)[

H(

D′′y |t)−H

((⌈D′′y∆′i

⌉−1)

∆′i|t)]

+∫ ⌈D′′

y∆′i

⌉∆′i

z=D′′/y η

(y⌈

D′′y∆′i

⌉∆′i,yz,D′′,D′

)dH (z|t)

yk1 (∆′i)

2(⌈

D′′y∆′i

⌉− D′′

y∆′i

)[D′′y∆′i−(⌈

D′′y∆′i

⌉−1)]

+yk1 (∆′i)

2(⌈

D′′y∆′i

⌉− D′′

y∆′i

)2

= yk1(∆′i)2(⌈

D′′

y∆′i

⌉− D′′

y∆′i

),

by L-H. There are now two cases.

Case 1:⌊

D′y∆′i

+1⌋≤ 1

∆′i. The last summand in (34), which corresponds to c= bD′/y∆′i +1c,

is3

∫ ⌊ D′y∆′i

+1⌋

∆′i

z=(⌊

D′y∆′i

+1⌋−1)

∆′i

η

(y⌊

D′

y∆′i+1⌋

∆′i,yz,D′′,D′

)dH (z|t)

=∫ D′/y

z=(⌊

D′y∆′i

+1⌋−1)

∆′i

η

(y⌊

D′

y∆′i+1⌋

∆′i,yz,D′′,D′

)dH (z|t)

(y⌊

D′

y∆′i+1⌋

∆′i,D′,D′′,D′

)[H(⌊

D′

y∆′i+1⌋

∆′i|t)−H

(D′

y|t)]

(35)

≤ yk1(∆′i)2[

D′

y∆′i+1−

⌊D′

y∆′i+1⌋]

. (36)

The remainder of the sum in (34) is⌊D′y∆′i

+1⌋−1

c=⌈

D′′y∆′i

⌉+1

∫ c∆′i

z=(c−1)∆′iη(yc∆′i,yz,D′′,D′

)dH (z|t)≤ yk1

(∆′i)2(⌊

D′

y∆′i+1⌋−⌈

D′′

y∆′i

⌉−1).

Collecting terms,∣∣ΩHyi (t, t,D′,D′′)

∣∣ ≤ k1∆′i (D′−D′′). Now take i∗ large enough that

k1∆′i∗ < ε .

3By Remark 1, line (35) is zero as y⌊

D′y∆′i

+1⌋

∆′i−D′ = y∆′i

(⌊D′y∆′i

+1⌋− D′

y∆′i

)> 0. The inequality in line

(36) then follows from Lipschitz-H.

28

Page 29:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Case 2:⌊

D′y∆′i

+1⌋> 1

∆′ior, equivalently,

⌊D′y∆′i

⌋> 1

∆′i−1. The final sum on the right hand

side of (34) then corresponds to c = 1/∆′i. Moreover,

1/∆′i

c=⌈

D′′y∆′i

⌉+1

∫ c∆′i

z=(c−1)∆′iη(yc∆′i,yz,D′′,D′

)dH (z|t)≤ yk1

(∆′i)2[

1∆′i−⌈

D′′

y∆′i

⌉−1].

Thus, ∣∣ΩHyi (t, t,D′,D′′)∣∣≤ yk1(∆′i)2[

1∆′i−1− D′′

y∆′i

]< yk1

(∆′i)2[⌊

D′

y∆′i

⌋− D′′

y∆′i

]≤ yk1

(∆′i)2[

D′

y∆′i− D′′

y∆′i

]= k1∆

′i(D′−D′′

)as before. Q.E.D.Claim 2

For all D,D′ ∈ [0,y] and all t ∈ Si\1, define the difference quotients of vHyi (D, t)

with respect to D and t:

∆Hyi1(D,D′, t

)=

vHyi (D, t)− vHyi (D′, t)D−D′

and (37)

∆Hyi2 (D, t) =

vHyi (D, t +∆i)− vHyi (D, t)∆i

. (38)

By parts 1 and 2 of Claim 2, if D < D′, then

∆Hyi1(D,D′, t

)∈(

k0

(1−∆

′i

⌈D′

y∆′i

⌉+∆

′i

),k1

(1− D

y+∆i

))⊂ (0,∞) (39)

while if D > y∆′i,

∆Hyi2 (D, t) ∈

(0,k1 min

D,y

(1−∆

′i))

. (40)

By (24), for any D ∈ [0,y], the partial derivatives of vHy (D, t) are given by

vHy2 (D, t) =−y

∫ D/y

z=0

∂H (z|t)∂ t

dz and (41)

vHy1 (D, t) = 1−H

(Dy|t)=∫ 1

z=D/y

∂H (z|t)∂ z

dz. (42)

For all D ∈ [0,y] and t ∈ [0,1],

∂vHy2 (D, t)∂D

=− ∂

∂ tH(

Dy|t)∈[

k0Dy

(1− D

y

),k1

)(43)

29

Page 30:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

by (41) and L-H; ∣∣∣∣∣∂vHy2 (D, t)

∂ t

∣∣∣∣∣≤ k2D (44)

by (41) and L-H;

∂vHy1 (D, t)∂D

=− ∂

∂DH(

Dy|t)∈(−k1

y,−k0

y

)(45)

by (42) and L-H;k0D2 [3y−2D]

6y2 < vHy2 (D, t)< k1D (46)

by (41) and L-H;

k0

(1− D

y

)< vHy

1 (D, t)< k1

(1− D

y

)(47)

by (42) and L-H; for all t ′ ∈ [0,1],∣∣∣vHy1 (D, t)− vHy

1(D, t ′

)∣∣∣= ∣∣∣∣H(Dy|t)−H

(Dy|t ′)∣∣∣∣≤ k1

∣∣t− t ′∣∣ (48)

by (42) and L-H; for all D′ ∈ [0,y],∣∣∣vHy1 (D, t)− vHy

1(D′, t

)∣∣∣= ∣∣∣∣H(Dy|t)−H

(D′

y|t)∣∣∣∣< k1

y

∣∣D−D′∣∣ (49)

by L-H. By (46) and (47),

0≤ k0D2 (3y−2D)

6yk1 (y−D)<

vHy2 (D, t)

vHy1 (D, t)

<k1yk0

(D

y−D

), (50)

and6yk1 (y−D)

k0D2 (3y−2D)>

vHy1 (D, t)

vHy2 (D, t)

>k0

k1y

(y−D

D

). (51)

By (43), (45), (46), (47), and (51),

∂D

(vHy

2 (D, t)

vHy1 (D, t)

)=

∂DvHy2 (D, t)

vHy1 (D, t)

−vHy

2 (D, t) ∂

∂DvHy1 (D, t)[

vHy1 (D, t)

]2 ∈[γ

y1 (D) ,γy

2 (D)]

(52)

where

γy1 (D) =

k0Dk1y

[1+

Dk0 (3y−2D)

6k1 (y−D)2

]≥ 0 (53)

30

Page 31:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

and

γy2 (D) =

k1yk0 (y−D)

[1+

Dk1

k0 (y−D)

]> 0. (54)

Note that γy1 (D) lies in (0,∞) if D ∈ (0,y), is zero if D = 0, and is ∞ if D = y. Moreover,

γy2 (D) is increasing in D, lies in (0,∞) if D∈ [0,y), and is ∞ if D= y. Finally, for 0≤ a≤ b,

maxD∈[a,b]

γy2 (D)≤ k1y

k0 (y−b)

[1+

bk1

k0 (y−b)

], (55)

which is positive and, if b < y, finite.

For any real number `, let (`,∞] and [`,∞] denote the sets (`,∞)∪∞ and [`,∞)∪∞,

respectively. Recall that f Hδy (D, t) is defined in (26). For t ∈ [0,1] and a ∈ (0,y], and

k ∈ (0,∞], we define the following modification of CPHδy:

CONTINUOUS IVP WITH PARAMETERS H,δ ,y, t,a,k (CPHδytak ). The differential

equationdDHδy

tak

dt= max

f Hδy

(DHδy

tak (t) , t),−k

(56)

with DHδytak : [t,1]→ℜ, together with the initial value DHδy

tak (t) = a.

Clearly, any DHδy0y∞

that solves CPHδy0y∞

must also be a solution DHδy to CPHδy and vice-

versa.

Claim 3. Consider any t ∈ [0,1], H in H , a ∈ (0,y], and k ∈ (0,∞].

1. If either a < y or k < ∞ (or both), then there exists a unique solution to CPHδytak , which

is decreasing and differentiable in t and takes values in (0,a].

2. Let δ , δ ∈ (0,1) and H, H ∈H satisfy f Hδy≤ f Hδy, and let a′ ∈ (0,a] and k′ ∈ [k,∞].

Suppose there exist (possibly nonunique) solutions DHδytak and DHδy

ta′k′ to CPHδytak and

CPHδyta′k′ , respectively. Then DHδy

tak (t)≥ DHδyta′k′ (t) for all t ∈ [t,1].

3. If either a < y or k < ∞ (or both), then the function DHδytak is Lipschitz continuous in

t with Lipschitz constant mink,ka where

ka =k1ya

(1−δ )k0 (y−a). (57)

31

Page 32:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

4. If a < y, then for all t ∈ [t,1], DHδytyka

(t)−DHδyta∞ (t) ∈ [0,y−a].

PROOF OF CLAIM 3. Part 1. We first show that max

f Hδy (D, t) ,−k

is (a) continuous

in t ∈ [0,1] and (b) Lipschitz continuous in D ∈ [0,a]. Since max is a Lipschitz-continuous

function, it suffices to show that f Hδy (D, t) has properties (a) and (b) whenever f Hδy (D, t)≥

−k. If k = ∞, then D ≤ a < y. If instead k < ∞ and f Hδy (D, t) ≥ −k then by (50),k0D2(3y−2D)

6yk1(y−D) < (1−δ )k; rearranging, D2

y−D < 6yk1(1−δ )kk0(3y−2D) ≤

6k1(1−δ )kk0

(since D ≤ y). SinceD2

y−D is continuous and increasing in D and approaches ∞ as D ↑ y, there is a constant

a′ < y such that D ≤ a′ for any D and t satisfying f Hδy (D, t) ≥ −k. Collecting cases,

D ≤ b = maxa,a′ < y. Hence 1−H(

Dy |t)≥ 1−H

(by |t)> 0; as H is continuously

differentiable in t, f Hδy (D, t) is continuous in t and thus satisfies (a). And by (52), (53),

and (55),∣∣∣ ∂

∂D f Hδy (D, t)∣∣∣ is at most k1y

k0(y−b)(1−δ )

[1+ bk1

k0(y−b)

], whence f Hδy (D, t) satisfies

(b). By the Picard-Lindelof theorem, there thus exists a unique solution DHδytak to CPHδy

tak .

The solution DHδytak is differentiable in t since the right hand side of (56) is finite. Finally,

f Hδy (D, t) < 0 for all D ∈ (0,y] and f Hδy (0, t) = 0. Hence, DHδytak (t) is decreasing in t

until and unless it hits zero, where it remains. Thus, by (50), DHδytak (t)≤ a for all t ∈ [t,1].

Finally, DHδytak (t)> 0 by the following lemma. Let DHδy

tak (t) first reach its minimum value

of D≥ 0 at t = t > t.

Lemma 1. The minimum face value D is nonzero.

PROOF OF LEMMA 1: For t ∈ [t, t], the function DHδytak (t) has a strictly decreasing inverse

tHδytak that satisfies the following inverse problem:

INVERSE CONTINUOUS IVP WITH PARAMETERS H,y, t,a,k (ICPHδytak ). The

differential equation

dtHδytak

dD=−max

(1−δ )vHy

1

(D, tHδy

tak (D))

vHy2

(D, tHδy

tak (D)) , 1

k

(58)

with tHδytak : [D,a]→ [t, t], together with the terminal value tHδy

tak (a) = t.

32

Page 33:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

By (51),dtHδy

takdD ≤−(1−δ ) k0

k1y

(y−D

D

). Hence, as tHδy

tak (a) = t,

t = tHδytak (D) = tHδy

tak (a)−∫ a

D=D

dtHδytak

dDdD

≥ t +(1−δ )k0

k4y

∫ y

D=D

(y−D

D

)dD = t +

(1−δ )k0

k4F(

Dy

),

where F (x) denotes x− lnx− 1, which is finite and differentiable for all finite x > 0. F

is decreasing in x ∈ (0,1): F ′ (x) = 1− 1/x < 0. Thus, F(

Dy

)< k4

(1−δ )k0(t− t), whence

D > yF−1(

k4(1−δ )k0

(t− t))

. Moreover, F (1) = 0 and limx↓0 F (x) = ∞, so the inverse F−1

is decreasing in x ∈ (0,∞) and satisfies F−1 (0) = 1 and limx→∞ F−1 (x) = 0. Hence, as

t− t > 0, D > 0. Q.E.D.Lemma 1

Part 2. Suppose not. Since DHδytak (t) = a ≥ a′ = DHδy

ta′k′ (t) and both solutions are

continuous in t, there must be t ≤ t0 < t1 ≤ 1 such that DHδytak (t0) = DHδy

ta′k′ (t0) and, for all

t ∈ (t0, t1], DHδytak (t)< DHδy

ta′k′ (t). Hence, by (56),

0 < DHδyta′k′ (t1)−DHδy

tak (t1) =∫ t1

t=t0

max

f Hδy(

DHδyta′k′ (t) , t

),−k′

−max

f Hδy

(DHδy

tak (t) , t),−k

dt

which is impossible: since f Hδy (D, t) is decreasing in D and k′ ≥ k, the integrand is

nonpositive for all t in (t0, t1].

Part 3 follows from (50) and the fact that DHδytak ≤ a.

Part 4. By part 2, Γa (t)d= DHδy

tyka(t)−DHδy

ta∞ (t)≥ 0. By (26),

Γ′a (t) = max

[

f Hδy(

DHδytyka

(t) , t)− f Hδy

(DHδy

ta∞ (t) , t)]

,

−ka− f Hδy(

DHδyta∞ (t) , t

) .

Both entries in the max are nonpositive by (50), (26), and the fact that DHδyta∞ (t) ≤ a.

Accordingly, for all t ∈ [t,1],∣∣∣DHδy

tyka(t)−DHδy

ta∞ (t)∣∣∣≤ Γa (t) = y−a. Q.E.D.Claim 3

Before addressing the discrete case, we prove some useful bounds:

33

Page 34:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Claim 4. Let w,w′,ζ e,ζ e′ be in (0,y] and satisfy w′ ≥ w, ζ e′ ≥ ζ e, w > ζ e, and w′ > ζ e′.

Then

0≤ ∆Hyi2(ζ e′, t

)−∆

Hyi2 (ζ e, t)≤ k1

(ζ e′−ζ e

). (59)

Moreover, if minw,w′,ζ ,ζ ′> y∆′i then

0≥ ∆Hyi1(ζ e′,w′, t

)−∆

Hyi1 (ζ e,w, t)≥−k1

[maxw′−w,ζ e′−ζ e

y+∆

′i

](60)

and∆

Hyi2 (ζ e, t)

∆Hyi1 (ζ e,w, t)

0,k1ζ e

k0

(1−∆′i

⌈w

y∆′i

⌉+∆′i

)⊂ (0,∞) . (61)

PROOF OF CLAIM 4. By (28), for any ζ e ∈ [0,y] and t ∈ Si\1,

vHyi (ζ e, t +∆i)− vHyi (ζ e, t)

= y∫ ζ e

y

z=0

[H(

∆′i

⌊z∆′i

⌋|t)−H

(∆′i

⌊z∆′i

⌋|t +∆i

)]dz. (62)

As the integrand is nonnegative, (62) is nondecreasing in ζ e, so ∆Hyi2 (ζ ′, t)−∆

Hyi2 (ζ , t)≥ 0.

By L-H, for any z ∈ [0,y], H (z|t)−H (z|t +∆i) < k1∆i. Equation (59) then follows from

(38) and (62).

By (28),

∆Hyi1(ζ e′,w′, t

)−∆

Hyi1 (ζ e,w, t) =

vHyi (ζ e′, t)− vHyi (w′, t)ζ e′−w′

− vHyi (ζ e, t)− vHyi (w, t)ζ e−w

=y

w−ζ e

∫ wy

z= ζ ey

H(

∆′i

⌊z∆′i

⌋|t)

dz− yw′−ζ e′

∫ w′y

z= ζ e′y

H(

∆′i

⌊z∆′i

⌋|t)

dz.

Define the change of variables z′ = ζ

y +(

w−ζ

w′−ζ ′

)(z− ζ ′

y

). When z = ζ ′

y , z′ = ζ

y , and when

z = w′y , z′ = w

y . Moreover, dz = w′−ζ ′

w−ζdz′ and z = ζ ′

y +(

w′−ζ ′

w−ζ

)(z′− ζ

y

)which we denote

ψ (z′). So∫ w′

y

z= ζ ′y

H(

∆′i

⌊z∆′i

⌋|t)

dz =∫ w

y

z′= ζ

y

H(

∆′i

⌊ψ(z′)

∆′i

⌋|t)

w′−ζ ′

w−ζdz′. Renaming z′ to z and

simplifying,

∆Hyi1(ζ e′,w′, t

)−∆

Hyi1 (ζ e,w, t)=

yw−ζ e

∫ wy

z= ζ ey

[H(

∆′i

⌊z∆′i

⌋|t)−H

(∆′i

⌊ψ (z)

∆′i

⌋|t)]

dz.

(63)

34

Page 35:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

We can write

z−ψ (z) =1

w−ζ e

((z− ζ e′

y

)[w−ζ e]−

(z− ζ e

y

)[w′−ζ e′

]). (64)

As the right hand side is linear in z, it reaches its maximum and minimum at the endpoints

of the interval of integration. At the lower endpoint (at z = ζ

y ), the right hand side of (64)

equals ζ−ζ ′

y , while at the upper endpoint (at z= wy ), it equals w−w′

y . Thus,−w≤ z−ψ (z)≤

−w where w = y−1 minw′−w,ζ ′−ζ and w= y−1 maxw′−w,ζ ′−ζ. As w and w are

both nonnegative, z≤ ψ (z), which by (63) establishes the first inequality in (60). Finally,

by (63) and L-H,

∆Hyi1(ζ e′,w′, t

)−∆

Hyi1 (ζ e,w, t)

≥ yw−ζ e

∫ wy

z′= ζ ey

[H(

∆′i

⌊ψ (z)−w

∆′i

⌋|t)−H

(∆′i

⌊ψ (z)

∆′i

⌋|t)]

dz

≥ yk1

w−ζ e

∫ wy

z′= ζ ey

[∆′i

⌊ψ (z)−w

∆′i

⌋−∆

′i

⌊ψ (z)

∆′i

⌋]dz

≥− yk1

w−ζ e

∫ wy

z′= ζ ey

[w+∆

′i]

dz =−k1[w+∆

′i].

This establishes the second inequality in (60). Finally, (61) follows from (39) and (40).

Q.E.D.Claim 4

Claim 5. Fix H, y, and i. Define

φ (D, t) = vHyi (D, t +∆i)−δvHyi (D, t) . (65)

For any D ∈ (y∆′i,y] and any t ∈ Si there exists a unique solution D∗ = D∗ (D), which lies

in (y∆′i,D), to

φ (D∗, t) = (1−δ )vHyi (D, t) . (66)

Moreover, D∗ (D) is increasing in D and D−D∗ (D) is nondecreasing in D. Finally,

∂φ

∂D≥ k0 (1−δ )

(1− D

y

). (67)

PROOF OF CLAIM 5. We first show three properties.

35

Page 36:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

1. φ (y∆′i, t) < (1−δ )vHyi (D, t). Proof: by (28), for all t in Si, vHyi (y∆′i, t) = y∆′i.

Hence, φ (y∆′i, t) = (1−δ )y∆′i. Moreover, vHyi (D, t) is strictly increasing in D ∈

[0,y] by part 2 of Claim 2, so vHyi (D, t) > y∆′i. The result then follows from (65)

and (66).

2. φ (D, t)> (1−δ )vHyi (D, t). Proof: for D ∈ (y∆′i,y], vHyi (D, t) is increasing in t by

part 1 of Claim 2. The result then follows from (65) and (66).

3. φ is continuous and increasing in D ∈ [0,y]. Proof: vHyi (D, t) is continuous in D by

(22), so φ (D, t) also is continuous in D. By (28),

φ (D, t) = (1−δ )D− y

[∫ Dy

z=0

[H(

∆′i

⌊z∆′i

⌋|t +∆i

)−δH

(∆′i

⌊z∆′i

⌋|t)]

dz

].

By L-H, 1−H (z|t)> k0 (1− z) for all z∈ [0,1] (and thus, substituting z = 0, k0 < 1),

so letting z0 = ∆′i

⌊D

y∆′i

⌋≤ D

y ,

∂φ

∂D= 1−δ − [H (z0|t +∆i)−δH (z0|t)]≥ (1−δ ) [1−H (z0|t)]

≥ (1−δ )k0 (1− z0)≥ k0 (1−δ )

(1− D

y

).

This establishes the result as well as equation (67).

Facts 1-3 imply that for any D ∈ (y∆′i,y], there exists a unique D∗ = D∗ (D) satisfying (66),

and that it lies in (y∆′i,D). Moreover, D∗ (D) is increasing in D.

Finally, let D0 > D and let D∗ = D∗(

D0

). To show that D−D∗ (D) is nondecreasing

in D, we must show that D0− D∗ ≥ D−D∗ or, equivalently, that D0−D ≥ D∗−D∗. By

(65),

φ (D, t) = vHyi (D, t +∆i)−δvHyi (D, t) = (1−δ )vHyi (D, t)+ vHyi (D, t +∆i)− vHyi (D, t) .

Hence, by (38), (59), and (66), and since D∗ > D∗,

(1−δ )[vHyi

(D0, t

)− vHyi (D, t)

]= φ

(D∗, t

)−φ (D∗, t)

≥ (1−δ )[vHyi

(D∗, t

)− vHyi (D∗, t)

],

36

Page 37:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

whence by part 2 of Claim 2,

D∗−D∗

D0−D≤

vHyi(D0,t)−vHyi(D,t)

D0−D

vHyi(D∗,t)−vHyi(D∗,t)

D∗−D∗

,

which is in (0,1] by (37) and (60). Thus, D∗−D∗ ≤ D0−D as claimed. Q.E.D.Claim 5

For any real number `, let (`,∞] and [`,∞] denote the sets (`,∞)∪∞ and [`,∞)∪∞,

respectively. For any constants t ∈ Si, a ∈ (0,y], and k ∈ (0,∞], consider the following

initial value problem, where Sti denotes the set of types t ≥ t in Si:

DISCRETE IVP WITH PARAMETERS H,y, t,a,k (DPHδyitak ). The condition

DHδyitak (t +∆i) = max

DHδyi∗

tak (t +∆i) ,DHδyitak (t)− k∆i

(68)

with DHδyitak : St

i →ℜ, where DHδyi∗tak (t +∆i) is the (by Claim 5) unique solution

D∗ ∈(

y∆′i,DHδyitak (t)

)to

vHyi (D∗, t +∆i)−δvHyi (D∗, t) = (1−δ )vHyi(

DHyitak (t) , t

), (69)

together with the initial value DHδyitak (t) = a > y∆′i.

Clearly, any DHδyi0y∞

that solves DPHδyi0y∞

must also be a solution DHδyi to DPHδyi and

vice-versa.

Claim 6. For any t ∈ Si, a ∈ (y∆′i,y], and k ∈ (0,∞]:

1. There exists a unique solution DHδyitak to DPHδyi

tak . This function is decreasing in t ∈ Sti

and takes values in (y∆′i,a].

2. Let a′ ∈ (y∆′i,a] and k′ ∈ [k,∞]. Then DHδyitak (t)≥ DHδyi

ta′k′ (t) for all types t ∈ Sti .

3. Let kia =

k1y(y+a)k0(1−δ )(y(1−∆′i)−a)

. For all types t ∈ Sti , 0 > DHδyi∗

tak (t +∆i)−DHδyitak (t) ≥

−∆ika and hence 0 > DHδyitak (t +∆i)−DHδyi

tak (t)≥−∆i mink,ka.

4. For all t ∈ Sti , DHδyi

tyka(t)−DHδyi

ta∞ (t) ∈ [0,y−a].

37

Page 38:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

PROOF OF CLAIM 6. Part 1. Follows from Claim 5.

Part 2. Clearly, DHδyitak (t)−DHδyi

ta′k′ (t) = a− a′ ≥ 0. And if, for some t ∈ Sti\1,

DHδyitak (t)≥DHδyi

ta′k′ (t), then DHδyitak (t)−k∆i≥DHδyi

ta′k′ (t)−k′∆i and, by Claim 5, DHδyi∗tak (t +∆i)≥

DHδyi∗ta′k′ (t +∆i), so DHδyi

tak (t +∆i)≥ DHδyita′k′ (t +∆i).

Part 3. Let D′ = DHδyitak (t), D′′ = DHδyi

tak (t +∆i), and D∗ = DHδyi∗tak (t +∆i). By Claim

5, D∗−D′ < 0 and minD′,D′′,D∗> y∆′i. We will show that D∗−D′ ≥−∆ika which, by

(68), implies 0 > D′′−D′ ≥−∆i mink,ka. The result is trivial when a = y since ky = ∞.

Suppose a < y. By (38), (40), (65), and (66),

φ(D′, t

)−φ (D∗, t) = φ

(D′, t

)− (1−δ )vHyi (D′, t)

= vHyi (D′, t +∆i)− vHyi (D′, t)≤ k1D′∆i ≤ k1a∆i.

But by (67), φ (D′, t)− φ (D∗, t) ≥ [D′−D∗]k0 (1−δ )(

1− ay

). Combining the two

inequalities and using (57) yields the result.

Part 4. Fix t ∈ [0,1] and a ∈ (0,y), whence ka ∈ (0,∞). By part 2, DHδyityka

(t) ≥

DHδyita∞ (t) for all types t ∈ St

i . Let Γa (t) = DHyityka

(t)−DHδyita∞ (t) ≥ 0. As DHδyi

ta∞ (t +∆i) =

DHδyi∗ta∞ (t +∆i),

Γa (t +∆i)−Γa (t)=max

DHδyi∗tyka

(t +∆i)−DHδyityka

(t)−[DHδyi∗

ta∞ (t +∆i)−DHδyita∞ (t)

],

−ka∆i−[DHδyi∗

ta∞ (t +∆i)−DHδyita∞ (t)

]

for any t < 1 in Sti by (68). By Claim 5, D′−D∗ (D′) is nondecreasing in D′, so

DHδyi∗tyka

(t +∆i)−DHδyityka

(t)−[DHδyi∗

ta∞ (t +∆i)−DHδyita∞ (t)

]≤ 0,

and by part 3, DHδyi∗ta∞ (t +∆i)−DHδyi

ta∞ (t) ≥ −∆ika. Hence, Γa (t +∆i) ∈ [0,Γa (t)], so for

all t ∈ Sti ,∣∣∣DHδyi

tyka(t)−DHδyi

ta∞ (t)∣∣∣≤ Γa (t) = y−a. Q.E.D.Claim 6

As already noted, we extend any function defined on Si to any t ∈ [0,1] by evaluating it

at τ it = ∆i bt/∆ic.

Claim 7. Fix t ∈ [0,1), a ∈ (y∆′i,y], and k ∈ (0,∞] such that either a < y or k < ∞ (or both).

For any a0,a1 ∈ (0,a] and any type t ∈ [t,1], DHδyita0k (t)−DHδy

ta1k (t) converges to zero as i→∞

and |a0−a1| → 0 (in either order), uniformly in H ∈H , y ∈ (0,y], and t.

38

Page 39:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

PROOF OF CLAIM 7. By (69), for any t ∈ [0,1],

(1−δ )∆Hyi1

(DHδyi∗

ta0k (t +∆i) ,DHδyita0k (t) ,τ i

t

)∆DHδyi∗

ta0k (t)+∆Hyi2

(DHδyi∗

ta0k (t +∆i) ,τit

)= 0

(70)

where we define

∆DHδyi∗ta0k (t) =

DHδyi∗ta0k (t +∆i)−DHδyi

ta0k (t)

∆i. (71)

Equation (70) can be rewritten

∆DHδyi∗ta0k (t) =− 1

1−δ

∆Hyi2

(DHδyi∗

ta0k (t +∆i) ,τit

)∆

Hyi1

(DHδyi∗

ta0k (t +∆i) ,DHδyita0k (t) ,τ i

t

) ,whence DHδyi∗

ta0k (t +∆i) = DHδyita0k (t)− 1

1−δ

∆Hyi2

(DHδyi∗

ta0k (t+∆i),τit

)∆

Hyi1

(DHδyi∗

ta0k (t+∆i),DHδyita0k (t),τ

it

)∆i and thus, by (68),

DHδyita0k (t +∆i) = DHδyi

ta0k (t)−∆i min

11−δ

∆Hyi2

(DHδyi∗

ta0k (t +∆i) ,τit

)∆

Hyi1

(DHδyi∗

ta0k (t +∆i) ,DHδyita0k (t) ,τ i

t

) ,k ,

and so

∆DHδyita0k (t) d

=DHδyi

ta0k (t +∆i)−DHδyita0k (t)

∆i

=−min

11−δ

∆Hyi2

(DHδyi∗

ta0k (t +∆i) ,τit

)∆

Hyi1

(DHδyi∗

ta0k (t +∆i) ,DHδyita0k (t) ,τ i

t

) ,k .

Lemma 2. In the above formula for ∆DHδyita0k (t), we may replace DHδyi∗

ta0k by DHδyita0k . That is,

∆DHδyita0k (t) =−min

11−δ

∆Hyi2

(DHδyi

ta0k (t +∆i) ,τit

)∆

Hyi1

(DHδyi

ta0k (t +∆i) ,DHδyita0k (t) ,τ i

t

) ,k . (72)

PROOF OF LEMMA 2. Let w = w′ = DHδyita0k (t). Let ζ e = DHδyi∗

ta0k (t +∆i) and ζ e′ =

DHδyita0k (t +∆i). By Claim 5 and part 1 of Claim 6, minw,ζ ,w′,ζ ′ > y∆′i. Hence, by

Claim 4, ∆Hyi1(ζ ′,w′,τ i

t)≤ ∆

Hyi1(ζ ,w,τ i

t)

and ∆Hyi2(ζ ′,τ i

t)≥ ∆

Hyi2(ζ ,τ i

t). Thus,

∆Hyi2(ζ e′,τ i

t)

∆Hyi1(ζ e′,w′,τ i

t) ≥ ∆

Hyi2(ζ e,τ i

t)

∆Hyi1(ζ e,w,τ i

t) , (73)

39

Page 40:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

with equality when ζ e′ = ζ e. Accordingly, there are two cases. If 11−δ

∆Hyi2 (ζ ,τ i

t )∆

Hyi1 (ζ ,w,τ i

t )≤ k,

then ζ e′ = ζ e, which implies (72). If 11−δ

∆Hyi2 (ζ ,τ i

t )∆

Hyi1 (ζ ,w,τ i

t )≥ k, then ∆DHδyi

ta0k (t) = −k but by

(73), 11−δ

∆Hyi2 (ζ ′,τ i

t )∆

Hyi1 (ζ ′,w′,τ i

t )≥ k as well, so (72) holds. Q.E.D.Lemma 2

For all t ′ ∈ [t,1], since DHδyita0k (t) = a0,

DHδyita0k

(t ′)= a0 +∆i ∑

t∈Sti ,t<τ i

t′

∆DHδyita0k (t)

= a0−∆i ∑t∈St

i ,t<τ it′

min

11−δ

∆Hyi2

(DHδyi

ta0k (t +∆i) , t)

∆Hyi1

(DHδyi

ta0k (t +∆i) ,DHδyita0k (t) , t

) ,k

= a0−∫

τ it′

t=tmin

11−δ

∆Hyi2

(DHδyi

ta0k (t +∆i) ,τit

)∆

Hyi1

(DHδyi

ta0k (t +∆i) ,DHδyita0k (t) ,τ i

t

) ,kdt, (74)

By (56) and (26), for all t ′ ∈ [t,1],

DHδyta1k

(t ′)= a1−

∫ t ′

t=tmin

11−δ

vt

(DHδy

ta1k (t) , t)

vD

(DHδy

ta1k (t) , t) ,kdt. (75)

Define the analogous quantities to ∆Hyi1 and ∆

Hyi2 with vHyi replaced by vHy: for any

D′,D′′ ∈ [0,y], let

∆Hy1(D′,D′′, t

)=

vHy (D′, t)− vHy (D′′, t)D′−D′′

∈(

k0

(1− maxD′,D′′

y

),k1

(1− minD′,D′′

y

))and (76)

∆Hy2(D′, t,∆i

)=

vHy (D′, t +∆i)− vHy (D′, t)∆i

(k0 (D′)

2 [3y−2D′]6y2 ,k1D′

), (77)

where the bounds follow from (46) and (47) and imply that

∆Hy2 (D′, t,∆i)

∆Hy1 (D′,D′′, t)

(k0 (D′)

2 [3y−2D′]6yk1 (y−minD′,D′′)

,yk1D′

k0 (y−maxD′,D′′)

). (78)

If a < y then by (61), (78), (50), and (57), for all D,D′ ∈ (y∆′i,a] and all t ∈ [0,1], the ratios∆

Hyi2 (D′,τ i

t )∆

Hyi1 (D′,D′′,τ i

t ), ∆

Hy2 (D′,t,∆i)

∆Hy1 (D′,t,τ i

t ),

vHyt (D′,τ i

t )vHy

D (D′,τ it )

, and vHyt (D′,t)

vHyD (D′,t)

are all at most (1−δ )ka. Let

κ = (1−δ )mink,ka< ∞. (79)

40

Page 41:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

It follows that by (74), (75) and the triangle inequality that for all t ′ ∈ [t,1],

(1−δ )∣∣∣DHδyi

ta0k

(t ′)−DHδy

ta1k

(t ′)∣∣∣=

∣∣∣∣∣∣∣∣∣∣∫ τ i

t′t=t min

Hyi2

(DHδyi

ta0k (t+∆i),τit

)∆

Hyi1

(DHδyi

ta0k (t+∆i),DHδyita0k (t),τ

it

) ,κ

dt

−∫ t ′

t=t min

vHy

t

(DHδy

ta1k(t),t)

vHyD

(DHδy

ta1k(t),t) ,κ

dt +a1−a0

∣∣∣∣∣∣∣∣∣∣≤ |a1−a0|+A1 +A2 +A3 +A4 +A5 +A6

where

A1 =

∣∣∣∣∣∣∫ t ′

t=τ it′

min

∆Hyi2

(DHδyi

ta0k (t +∆i) ,τit

)∆

Hyi1

(DHδyi

ta0k (t +∆i) ,DHδyita0k (t) ,τ i

t

) ,κdt

∣∣∣∣∣∣ ,

A2 =∫ t ′

t=t

∣∣∣∣∣∣∣∣∣∣min

Hyi2

(DHδyi

ta0k (t+∆i),τit

)∆

Hyi1

(DHδyi

ta0k (t+∆i),DHδyita0k (t),τ

it

) ,κ

−min

Hy2

(DHδyi

ta0k (t+∆i),τit ,∆i

)∆

Hy1

(DHδyi

ta0k (t+∆i),DHδyita0k (t),τ

it

) ,κ∣∣∣∣∣∣∣∣∣∣dt,

A3 =∫ t ′

t=t

∣∣∣∣∣∣∣∣∣∣min

Hy2

(DHδyi

ta0k (t+∆i),τit ,∆i

)∆

Hy1

(DHδyi

ta0k (t+∆i),DHδyita0k (t),τ

it

) ,κ

−min

vHy

t

(DHδyi

ta0k (t+∆i),τit

)vHy

D

(DHδyi

ta0k (t),τit

) ,κ

∣∣∣∣∣∣∣∣∣∣dt,

A4 =∫ t ′

t=t

∣∣∣∣∣∣min

vHyt

(DHδyi

ta0k (t +∆i) ,τit

)vHy

D

(DHδyi

ta0k (t) ,τ it

) ,κ

−min

vHyt

(DHδyi

ta0k (t) ,τ it

)vHy

D

(DHδyi

ta0k (t) ,τ it

) ,κ∣∣∣∣∣∣dt,

A5 =∫ t ′

t=t

∣∣∣∣∣∣min

vHyt

(DHδyi

ta0k (t) ,τ it

)vHy

D

(DHδyi

ta0k (t) ,τ it

) ,κ−min

vHyt

(DHδy

ta1k (t) ,τit

)vHy

D

(DHδy

ta1k (t) ,τit

) ,κ∣∣∣∣∣∣dt,

and

A6 =∫ t ′

t=t

∣∣∣∣∣∣min

vHyt

(DHδy

ta1k (t) ,τit

)vHy

D

(DHδy

ta1k (t) ,τit

) ,κ−min

vHyt

(DHδy

ta1k (t) , t)

vHyD

(DHδy

ta1k (t) , t) ,κ

∣∣∣∣∣∣dt.

Clearly, A1 ≤ κ∆i. For A2, A3, A4, and A5, we require the following claim.

41

Page 42:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Lemma 3. For any a,b,c,d ≥ 0, max|a−b| , |c−d| is an upper bound on both

|mina,c−minb,d|

and |maxa,c−maxb,d|.

PROOF OF LEMMA 3. We prove the result for |mina,c−minb,d|; the proof for

|maxa,c−maxb,d| is identical with the keyword min replaced max throughout. First,

assume a ≥ b and c ≥ d. Then |mina,c−minb,d| = mina,c−minb,d. And

max|a−b| , |c−d|= maxa−b,c−d. But

mina,c−minb,d ≤maxa−b,c−d

since

mina,c ≤minmaxa,b+ c−d ,maxc,d +a−b

= minb+maxa−b,c−d ,d +maxa−b,c−d

= minb,d+maxa−b,c−d .

The other cases (in which b > a or c > d or both) are analogous. Q.E.D.Lemma 3

This Lemma leads to the following useful bound.

Lemma 4. For any a,b,κ ∈ (0,∞) and c,d ∈ [0,∞),∣∣∣∣mina

c,κ−min

bd,κ

∣∣∣∣≤ ∣∣∣∣ a−bmaxc,a/κ

∣∣∣∣+ bmaxd,b/κ

max|c−d| , |a−b|/κmaxc,a/κ

.

(80)

PROOF OF LEMMA 4. For any a,b≥ 0 and c,d > 0,∣∣∣∣ac − bd

∣∣∣∣= ∣∣∣∣ad−bccd

∣∣∣∣≤ ∣∣∣∣ad−bdcd

∣∣∣∣+ ∣∣∣∣bd−bccd

∣∣∣∣= ∣∣∣∣a−bc

∣∣∣∣+ bd

∣∣∣∣d− cc

∣∣∣∣ . (81)

Moreover, for any a > 0 and c,κ ≥ 0,

mina

c,κ= κ min

acκ

,1= κ min

acκ

,aa

=

maxcκ,a=

amaxc,a/κ

. (82)

42

Page 43:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

By (81) and (82) and using Lemma (in that order),∣∣∣∣mina

c,κ−min

bd,κ

∣∣∣∣= ∣∣∣∣ amaxc,a/κ

− bmaxd,b/κ

∣∣∣∣≤∣∣∣∣ a−bmaxc,a/κ

∣∣∣∣+ bmaxd,b/κ

∣∣∣∣maxd,b/κ−maxc,a/κmaxc,a/κ

∣∣∣∣≤∣∣∣∣ a−bmaxc,a/κ

∣∣∣∣+ bmaxd,b/κ

max|c−d| , |a−b|/κmaxc,a/κ

.

Q.E.D.Lemma 4

Let D′t = DHδyita0k (t +∆i) ∈ (y∆′i,a0), D′′t = DHδyi

ta0k (t) ∈ (y∆′i,a0], at = ∆Hyi2(D′t ,τ

it), bt =

∆Hy2(D′t ,τ

it ,∆i

), ct = ∆

Hyi1(D′t ,D

′′t ,τ

it), and dt = ∆

Hy1(D′t ,D

′′t ,τ

it). By (80),

A2 ≤∫ t ′

t=t

∣∣∣∣ at−bt

maxct ,at/κ

∣∣∣∣dt +∫ t ′

t=t

(bt

maxdt ,bt/κmax|ct−dt | , |at−bt |/κ

maxct ,at/κ

)dt,

By (37), (38), (76), (77), and parts 3 and 4 of Claim 2, for any ε > 0 there is an i∗ < ∞

such that if i > i∗, max|at−bt | , |ct−dt |< ε . By (77), bt ∈[

k0(D′t)2

6y ,k1y]

. By part 1 of

Claim 2, at > k0(D′)2(3y−2D′)

6y2 > k0(D′)2

6y where D′ = D′t − 2y∆′i < y. Hence, minat ,bt >k0(D′t−2y∆′i)

2

6y . By (37), (76), and part 2 of Claim 2, minct ,dt ≥ k0

(1− D′′t

y

)so since

D′t ≥ D′′t − k∆i, both maxct ,at/κ and maxdt ,bt/κ are at least

k0 max

1− D′′t

y,(D′t−2y∆′i)

2

6yκ

≥ k0 max

1− D′′t

y,(D′′t − k∆i−2y∆′i)

2

6yκ

,

which is bounded below by a strictly positive constant κ ′ for large enough i as y > 0.

Collecting these bounds, A2 ≤ εκ2 [t ′− t] where κ2 =1κ ′

[1+ k1y

κ ′ max1,1/κ]∈ (0,∞).

By (80), redefining at = ∆Hy2(D′t ,τ

it ,∆i

), bt = vHy

t(D′t ,τ

it), ct = ∆

Hy1(D′t ,D

′′t ,τ

it), and

dt = vHyD(D′′t ,τ

it),

A3 ≤∫ t ′

t=t

∣∣∣∣ at−bt

maxct ,at/κ

∣∣∣∣dt +∫ t ′

t=t

(bt

maxdt ,bt/κmax|ct−dt | , |at−bt |/κ

maxct ,at/κ

)dt.

By (46), bt ≤ k1y. By the Mean Value Theorem, there is a t ∈[τ i

t ,τit +∆i

]such that

vHyt (D′t , t) = at . Thus, by (44),

|at−bt |=∣∣∣vHy

t(D′t , t

)−bt

∣∣∣= ∣∣∣vHyt(D′t , t

)− vHy

t(D′t ,τ

it)∣∣∣≤ k2y∆i.

43

Page 44:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Also by the Mean Value Theorem, there is a D ∈ [D′t ,D′′t ] such that vHy

D(D,τ i

t)= ct . By

part 3 of Claim 6 and (79),

∣∣D′′t −D′t∣∣≤ ∆i mink,ka=

κ

(1−δ )∆i. (83)

Hence, |ct−dt | =∣∣∣vHy

D(D,τ i

t)− vHy

D(D′′t ,τ

it)∣∣∣ ≤ k1

y |D′′t −D′t | ≤

k1κ

y(1−δ )∆i. By (46), (47),

(76), and (77), minat ,bt ≥ k0(D′t)2

6y , and minct ,dt ≥ k0

(1− D′′t

y

)so as shown in the

prior paragraph, both maxct ,at/κ and maxdt ,bt/κ are at least κ ′ > 0. Collecting

these bounds, A3 ≤ ∆iκ3 [t ′− t] where κ3 =yκ ′

[k2 +

k1κ ′ max

k1κ

y(1−δ ) ,k2yκ

]∈ (0,∞).

By (80), redefining at = vHyt(D′t ,τ

it), bt = vHy

t(D′′t ,τ

it), and ct = dt = vHy

D(D′′t ,τ

it),

A4 ≤∫ t ′

t=t

∣∣∣∣ at−bt

maxct ,at/κ

∣∣∣∣dt +∫ t ′

t=t

(bt

maxdt ,bt/κ|at−bt |/κ

maxct ,at/κ

)dt,

By (46), bt ≤ k1y. By (49) and (83),

|at−bt |=∣∣∣vHy

t(D′t ,τ

it)− vHy

t(D′′t ,τ

it)∣∣∣≤ k1κ

(1−δ )y∆i.

By (46) and (47), minat ,bt ≥ k0(D′t)2

6y , and minct ,dt ≥ k0

(1− D′′t

y

)so as shown in the

prior paragraph, both maxct ,at/κ and maxdt ,bt/κ are at least κ ′ > 0. Collecting

these bounds, A4 ≤ ∆iκ4 [t ′− t], where κ4 =k1

κ ′(1−δ )

y +k1κ ′

]∈ (0,∞).

By (80), redefining at = vHyt(D′′t ,τ

it), bt = vHy

t

(DHδy

ta1k

(τ i

t),τ i

t

), ct = vHy

D(D′′t ,τ

it), and

dt = vHyD

(DHδy

ta1k

(τ i

t),τ i

t

),

A5 ≤∫ t ′

t=t

∣∣∣∣ at−bt

maxct ,at/κ

∣∣∣∣dt +∫ t ′

t=t

(bt

maxdt ,bt/κmax|ct−dt | , |at−bt |/κ

maxct ,at/κ

)dt.

By (46), bt ≤ k1y. By (44), |at−bt | ≤ k2y∣∣∣D′′t −DHδy

ta1k

(τ i

t)∣∣∣. By (49),

|ct−dt | ≤k1

y

∣∣∣D′′t −DHδyta1k

it)∣∣∣ .

By (46), (47), minat ,bt≥ k0(D′t)2

6y and minct ,dt≥ k0

(1− D′′t

y

)so as shown above, both

maxct ,at/κ and maxdt ,bt/κ are at least κ ′ > 0. Collecting these bounds and using

44

Page 45:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

DHδyita0k (t) = DHδyi

ta0k

(τ i

t),

A5 ≤ κ5

∫ t ′

t=t

∣∣∣DHδyita0k

it)−DHδy

ta1k

it)∣∣∣dt ≤ κ5

[t ′− t

]maxt∈[t,t ′]

∣∣∣DHδyita0k

it)−DHδy

ta1k

it)∣∣∣

≤ κ5[t ′− t

]maxt∈[t,t ′]

∣∣∣DHδyita0k (t)−DHδy

ta1k (t)∣∣∣ ,

where κ5 =yκ ′

[k2 +

k1κ ′ max

k1y ,

k2yκ

]∈ (0,∞).

Now redefine at = vHyt

(DHδy

ta1k (t) ,τit

), bt = vHy

t

(DHδy

ta1k (t) , t)

, ct = vHyD

(DHδy

ta1k (t) ,τit

),

and dt = vHyD

(DHδy

ta1k (t) , t)

. By (80),

A6 ≤∫ t ′

t=t

∣∣∣∣ at−bt

maxct ,at/κ

∣∣∣∣dt +∫ t ′

t=t

(bt

maxdt ,bt/κmax|ct−dt | , |at−bt |/κ

maxct ,at/κ

)dt.

By (46), bt ≤ k1y. By (44), |at−bt | ≤ k2y∆i. By (48), |ct−dt | ≤ k1∆i. By (46)

and (47), minat ,bt ≥ k0(D′t)2

6y and minct ,dt ≥ k0

(1− D′′t

y

)so as shown above, both

maxct ,at/κ and maxdt ,bt/κ are at least κ ′ > 0. Collecting these bounds, A6 ≤

∆iκ6 [t ′− t] where κ6 =yκ ′

[k2 +

k1κ ′ max

k1,

k2yκ

]∈ (0,∞).

Summarizing our findings and since limi→∞ ∆i = 0 and t ′ ∈ [t,1], for all ε > 0 there is

an i∗ < ∞ such that if i > i∗, then

(1−δ )∣∣∣DHδyi

ta0k

(t ′)−DHδy

ta1k

(t ′)∣∣∣≤ |a1−a0|+κ

′′ε +κ5

[t ′− t

]maxt∈[t,t ′]

∣∣∣DHδyita0k (t)−DHδy

ta1k (t)∣∣∣

where κ ′′ = κ +(κ2 +κ3 +κ4 +κ6) [1− t]. So for any t ′′ ∈ [t, t ′],

(1−δ )∣∣∣DHδyi

ta0k

(t ′′)−DHδy

ta1k

(t ′′)∣∣∣

≤ |a1−a0|+κ′′ε +κ5

[t ′′− t

]max

t∈[t,t ′′]

∣∣∣DHδyita0k (t)−DHδy

ta1k (t)∣∣∣

≤ |a1−a0|+κ′′ε +κ5

[t ′− t

]maxt∈[t,t ′]

∣∣∣DHδyita0k (t)−DHδy

ta1k (t)∣∣∣

and therefore,

(1−δ ) maxt∈[t,t ′]

∣∣∣DHδyita0k (t)−DHδy

ta1k (t)∣∣∣

≤ |a1−a0|+κ′′ε +κ5

[t ′− t

]maxt∈[t,t ′]

∣∣∣DHδyita0k (t)−DHδy

ta1k (t)∣∣∣ .

45

Page 46:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Now for t ′ ∈ [t, t +b] where b = 1−δ

2κ5> 0, (1−δ )−κ5 [t ′− t]≥ 1−δ

2 , so

maxt∈[t,t ′]

∣∣∣DHδyita0k (t)−DHδy

ta1k (t)∣∣∣≤ 2

1−δ

(|a1−a0|+κ

′′ε),

whence maxt∈[t,t+b]

∣∣∣DHδyita0k (t)−DHδy

ta1k (t)∣∣∣≤ 2

1−δ(|a1−a0|+κ ′′ε). In particular,∣∣∣DHδyi

ta0k (t +b)−DHδyta1k (t +b)

∣∣∣≤ 21−δ

(|a1−a0|+κ

′′ε).

Let a2 = DHδyita0k (t +b) and a3 = DHδy

ta1k (t +b). Since DHδyita0k and DHδy

ta1k are decreasing

functions, maxa2,a3 < a so in the above reasoning we can use the same constant a

and thus the same constant ka and thus the same κ ′′. Accordingly,

maxt∈[t+b,t+2b]

∣∣∣DHδyita0k (t)−DHδy

ta1k (t)∣∣∣≤ 2

1−δ

(|a3−a2|+κ

′′ε)

≤ 21−δ

(2

1−δ

(|a1−a0|+κ

′′ε)+κ

′′ε

)=

(2

1−δ

)2

|a1−a0|+

[2

1−δ+

(2

1−δ

)2]

κ′′ε.

Iterating this reasoning n =⌈

1−tb

⌉(which does not depend on i) times, we obtain

maxt∈[t,1]

∣∣∣DHδyita0k (t)−DHδy

ta1k (t)∣∣∣≤ ( 2

1−δ

)n

|a1−a0|+κ′′ε

n

∑i=1

[(2

1−δ

)i].

Since the constants multiplying |a1−a0| and ε are independent of i, t ∈ [t,1], y ∈ (0,y],

and H ∈H , the result follows. Q.E.D.Claim 7

Claim 8. For all t ∈ [0,1), there exist unique solutions DHδyty∞

and DHδyity∞

to CPHδyty∞

and

DPHδyity∞

, respectively. They are decreasing in t and, in the case of DHδyty∞

, continuous in

t ∈ [t,1].

PROOF OF CLAIM 8. Define, for all t ∈ [t,1],

DHδyt (t) = lim

a↑yDHδy

tyka(t) = lim

a↑yDHδy

ta∞ (t) , (84)

and for all t ∈ Sti , DHδyi

t (t) = lima↑y DHδyityka

(t) = lima↑y DHδyita∞ (t). The four limits exist by

part 2 of Claims 3 and 6 and the Monotone Convergence Theorem. The pair of limits that

46

Page 47:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

appears in each equation are equal by part 4 of Claims 3 and 6. Hence, DHδyt and DHδyi

t

exist and are unique. Moreover, by part 2 of Claims 3 and 6, any solutions DHδyty∞

and DHδyity∞

to CPHδyty∞

and DPHδyity∞

must satisfy DHδytyka≥ DHδy

ty∞≥ DHδy

ta∞ and DHδyityka≥ DHδyi

ty∞≥ DHδyi

ta∞ for

any a ∈ (0,y], and thus must coincide with DHδyt and DHδyi

t , respectively. It remains to

show that DHδyt and DHδyi

t solve CPHδyty∞

and DPHδyity∞

: that

DHδyt (t) = y and, for t ∈ (t,1) ,

dDHδyt (t)dt

= f Hδy(

DHδyt (t) , t

)(85)

and

DHδyit (t) = y and, for t ∈ St

i\1 , DHδyit (t +∆i) = DHδyi∗

t (t +∆i) (86)

where DHδyi∗t (t +∆i) is the (by Claim 5) unique solution D∗ ∈

(y∆′i,D

Hδyit (t)

)to

vHyi (D∗, t +∆i)−δvHyi (D∗, t) = (1−δ )vHyi(

DHδyit (t) , t

). (87)

First, DHδyt (t) = lima↑y DHδy

ta∞ (t) = lima↑y a = y. Second, for t ∈ (t,1) , we must show

that dDHδyt (t)/dt = f Hδy

(DHδy

t (t) , t)

. By (52) and (26), f Hδy (ζ , t) is continuous in t

so, by (56), f Hδy(

DHδyt (t) , t

)is the limit of the derivatives dDHδy

ta∞ (t)/dt as a goes to y.

We now invoke the following well known result.4

Theorem 2. Let ( fn)∞

n=1 be a sequence of real-valued functions on t ∈ [t,1]. Suppose that

they converge uniformly to some function f . Assume also that the sequence of derivatives

( f ′n)∞

n=1 converges, uniformly in t ∈ [t,1], to some continuous function. Then f is differentiable

and limn→∞ f ′n (t) = f ′ (t) for all t ∈ [t,1].

Let fn = DHδytan∞ where (an)

n=1 is an increasing sequence that converges to y. By part 4

of Claim 3, ( fn)∞

n=1 converges to f = DHδyt , uniformly in t ∈ [t,1]. Fix w ∈ (t,1). We will

show that the sequence of derivatives ( f ′n)∞

n=1 converges, uniformly in t ∈ [w,1], to f ′. The

result then follows by taking w→ t. By (50) and (26), letting b = 11−δ

k06k1

and x = fn (w),

x = fn (t)+∫ w

t=tf ′n (t)dt = an +

∫ w

t=tf Hδy ( fn (t) , t)dt < an−

by

∫ w

t=t[ fn (t)]

2 dt

< an−by

∫ w

t=tx2dt = an−

by

x2 (w− t)

4See, e.g., Theorem 9.13 in Apostol (1981, p. 229).

47

Page 48:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

so x lies below the higher root of bx2 (w− t)+yx−yan = 0, which (since an ≤ y) is at most−y+√

y2+4b(w−t)y2

2b(w−t) . If we set this equal to y(1− c) and solve for c, we obtain c(1−c)2 =

b(w− t). Hence, for all t ∈ [w,1],

fn (t)y≤ fn (w)

y< 1− c (88)

where c is positive and independent of n. Thus, for all t ∈ [w,1] and all n,

∣∣ f ′n (t)∣∣=− f Hδy ( fn (t) , t)<1

1−δ

k1yk0

1− cc

which is a finite constant that is independent of n. Thus, the functions ( fn)∞

n=1, as well as

their limit (call it f∞), are Lipschitz on t ∈ [w,1] with the same Lipschitz constant. As f Hδy

is continuous and f∞ is Lipschitz continuous, the function

limn→∞

f ′n (t) = limn→∞

f Hδy ( fn (t) , t) = f Hδy(

limn→∞

fn (t) , t)= f Hδy ( f∞ (t) , t)

is continuous in t ∈ [w,1]. Moreover, convergence of f ′n (t) is uniform since, for all n,n′≥ 1,

by (52), (26), and (88), letting k5 =k1

k0c(1−δ )

[1+ k1(1−c)

k0c

]∈ (0,∞),

∣∣ f ′n (t)− f ′n′ (t)∣∣= ∣∣∣ f Hδy ( fn (t) , t)− f Hδy ( fn′ (t) , t)

∣∣∣≤ k5 | fn (t)− fn′ (t)|

= k5

∣∣∣DHδytan∞ (t)−DHδy

tan′∞(t)∣∣∣≤ y−minan,an′ ,

where the last inequality is from part 4 of Claim 3. Hence, ( f ′n)∞

n=1 is a uniform (in

t ∈ [w,1]) Cauchy sequence and thus converges uniformly. This proves existence and

uniqueness.

By part 1 of claim 6, DHδyity∞

is decreasing in t. It remains to show that DHδyty∞

is decreasing

and continuous in t ∈ [t,1]. By L-H and (26), f Hδy (D, t) is finite and negative for all D< y.

Thus, by (85), DHδyt is decreasing and continuous in t ∈ (t,1]. By (85), for continuity

at t = t we must show that limt→t DHδyt (t) = y. By (84), DHδy

t (t) = lima↑y DHδytyka

(t) =

lima↑y DHδyta∞ (t). By CPHδy

tak , DHδytak (t) = a. By the triangle inequality, for any a ∈ (0,y),∣∣∣DHδy

t (t)− y∣∣∣= ∣∣∣DHδy

t (t)−DHδyta∞ (t)

∣∣∣+ ∣∣∣DHδyta∞ (t)−a

∣∣∣+ |a− y| .

48

Page 49:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

By parts 2 and 4 of Claim 3,∣∣∣DHδy

t (t)−DHδyta∞ (t)

∣∣∣≤ ∣∣∣DHδytyka

(t)−DHδyta∞ (t)

∣∣∣≤ |y−a|. Finally,∣∣∣DHδyta∞ (t)−a

∣∣∣= ∣∣∣DHδyta∞ (t)−DHδy

ta∞ (t)∣∣∣≤ k1ya

(1−δ )k0(y−a) |t− t| by part 3 of Claim 3. Collecting

terms and letting a = y−√

t− t ≤ y,∣∣∣DHδy

t (t)− y∣∣∣ ≤ [2+ k1y2

(1−δ )k0

]√t− t which goes to

zero as t ↓ t. Q.E.D.Claim 8

In light of Claim 8, part 5 of Theorem 1 is implied by part 2 of Claim 3, setting t = 0,

a = a′ = y, and k = k′ = ∞.

We now turn to the convergence of the discrete solutions to the continuous one. By

part 2 of Claims 3 and 6, for all a ∈ (0,y),

DHδyita∞ (t)−DHδy

tyka(t)≤ DHδyi

ty∞(t)−DHδy

ty∞(t)≤ DHδyi

tyka(t)−DHδy

ta∞ (t)

and thus, by the triangle inequality,

∣∣∣DHδyity∞

(t)−DHδyty∞

(t)∣∣∣≤max

∣∣∣DHδyi

ta∞ (t)−DHδyta∞ (t)

∣∣∣+ ∣∣∣DHδyta∞ (t)−DHδy

tyka(t)∣∣∣ ,∣∣∣DHδyi

tyka(t)−DHδy

tyka(t)∣∣∣+ ∣∣∣DHδy

tyka(t)−DHδy

ta∞ (t)∣∣∣ . (89)

Fix ε > 0. By Claim 7, there is an i∗, independent of y, H, and t, such that for i > i∗,∣∣∣DHδyita∞ (t)−DHδy

ta∞ (t)∣∣∣ and

∣∣∣DHδyityk (t)−DHδy

tyk (t)∣∣∣ are each less than ε/2. By part 4 of Claim

3, for all a ∈ [y− ε/2,y),∣∣∣DHδy

tyka(t)−DHδy

ta∞ (t)∣∣∣≤ ε/2. But (89) holds for all a ∈ (0,y), so

it holds in particular for a ∈ [y− ε/2,y). Thus, for all ε > 0 there is an i∗, independent of

y, H, and t, such that for i > i∗,∣∣∣DHδyi

ty∞(t)−DHδy

ty∞(t)∣∣∣≤ ε . We now set t = 0 to obtain the

desired result. Together with Claim 8, this proves that the unique solution DHδyity∞

to DPHδyity∞

converges to the unique solution DHδyty∞

to CPHδyty∞

as i→ ∞, uniformly in H ∈H , y ∈ [0,y],

and t ∈ [0,1].

We next prove that pHδy and thus uHδy is both continuous and decreasing in the type t.

First, pHδy (t) = vHy(

DHδyty∞

(t) , t)

. By (24) and L-H, vHy is continuous in both arguments.

Since DHδyty∞

is also continuous in t, so is pHδy. By (27),

ddt

[vHy(

DHδy (t) , t)]

= vHy1

(DHδy (t) , t

) dDHδy

dt+ vHy

2

(DHδy (t) , t

)= δvHy

1

(DHδy (t) , t

) dDHδy

dt

49

Page 50:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

which is negative by (27), (46), and (47). Hence, pHδy is decreasing in t.

We now turn to convergence of pHδyi to pHδy and thus of uHδyi to uHδy. For any

t ∈ [0,1], recall that τ it = ∆i bt/∆ic. As DHδyi (t) = DHδyi (τ i

t),∣∣∣pHδyi (t)− pHδy (t)

∣∣∣= ∣∣∣pHδyi (τ

it)− pHδy (t)

∣∣∣=∣∣∣vHyi

(DHδyi (t) ,τ i

t

)− vHy

(DHδy (t) , t

)∣∣∣≤ A′1 +A′2 +A′3

where

A′1 =∣∣∣vHyi

(DHδyi (t) ,τ i

t

)− vHy

(DHδyi (t) ,τ i

t

)∣∣∣ ,A′2 =

∣∣∣vHy(

DHδyi (t) ,τ it

)− vHy

(DHδy (t) ,τ i

t

)∣∣∣ , and

A′3 =∣∣∣vHy

(DHδy (t) ,τ i

t

)− vHy

(DHδy (t) , t

)∣∣∣ .By (24), (28), and since H

(0|τ i

t)

is zero,

A′1y≤∫ DHδyi(t)/y

z=0

∣∣∣∣H (z|τ it)−H

(∆′i

⌊z∆′i

⌋|τ i

t

)∣∣∣∣dz≤ k1∆′i,

where the second inequality is by L-H. By (24), A′2 ≤ 2∣∣∣DHδyi (t)−DHδy (t)

∣∣∣. By (24) and

L-H,

A′3 ≤ y∫ DHδy(t)/y

z=0

∣∣H (z|t)−H(z|τ i

t)∣∣dz≤ yk1

∣∣t− τit∣∣≤ yk1∆i.

Hence, by the prior result, for all ε > 0 there is an i∗ < ∞ such that if i > i∗,∣∣∣pHδyi (t)− p(t)∣∣∣≤ yk1∆

′i +2

∣∣∣DHδyi (t)−DHδy (t)∣∣∣+ yk1∆i < ε,

for all t ∈ [0,1], y ∈ (0,y], and H ∈H , as claimed. Hence, pHδyi converges uniformly to

pHδy.

We now show uniform convergence of ΠHδyi to ΠHδy:∣∣∣ΠHδyi (t)−ΠHδy (t)

∣∣∣≤ ∣∣E i [yZ|t]−E∞ [yZ|t]∣∣+ ∣∣∣uHδyi (t)−uHδy (t)

∣∣∣

50

Page 51:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

and for z ∈ [(c−1)∆′i,c∆′i), c =⌊

z∆′i

⌋+1, so

∣∣E i [yZ|t]−E∞ [yZ|t]∣∣= ∣∣∣∣∣1/∆′i

∑c=1

yc∆′i[H(c∆′i|t)−H

((c−1)∆

′i|t)]−∫ 1

z=0yzdH (z|t)

∣∣∣∣∣= y

∣∣∣∣∣1/∆′i

∑c=1

∫ c∆′i

z=(c−1)∆′i

(⌊z∆′i

⌋+1)

∆′idH (z|t)−

∫ 1

z=0zdH (z|t)

∣∣∣∣∣≤ y∆

′i

∫ 1

z=0

∣∣∣∣⌊ z∆′i

⌋+1− z

∆′i

∣∣∣∣dH (z|t)≤ y∆′i. (90)

Thus, E i [yZ|t] converges uniformly to E∞ [yZ|t] and hence ΠHδyi converges uniformly to

ΠHδy.

As there is no mention of G in the statement of the problems DPHδyi and CPHδy, any

solutions DHδyi and DHδy to these problems for one distribution G are also solutions for

any other distribution G that satisfies our assumptions.5 Since, moreover, the solutions

DHδyi and DHδy are unique by parts 1 and 2 of the theorem, they must be independent of

G. Hence, convergence of DHδyi, pHδyi, uHδyi, and ΠHδyi is uniform in G as well.

We now show that EuGHδyi converges uniformly to EuGHδy. Since G has no atoms,

G(0) = 0; hence,

EuGHδyi =1/∆i

∑c=1

uHδyi (c∆i) [G(c∆i)−G((c−1)∆i)] =∫ 1

t=0uHδyi (

τit +∆i

)dG(t)

and thus, EuGHδyi−EuGHδy = A′′1 +A′′2−A′′3 where

A′′1 =∫ 1

t=0

[uHδyi (

τit +∆i

)−uHδy (

τit +∆i

)]dG(t) ,

A′′2 =∫ 1

t=0 uHδy (τ it +∆i

)dG(t), and A′′3 =

∫ 1t=0 uHδy (t)dG(t). As shown above, for all

ε > 0 there is an i∗ such that for all models i > i∗, parameters y in (0,y], and conditional

distribution functions H in H ,

∣∣A′′1∣∣≤ ∫ 1

t=0

∣∣∣uHδyi (τ

it +∆i

)−uHδy (

τit +∆i

)∣∣∣dG(t)<∫ 1

t=0

ε

2dG(t) =

ε

2.

5In particular, G must have support [0,1].

51

Page 52:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Moreover, since uHδy is a nonincreasing function of t, it follows that

uHδy (τ

it +∆i

)≤ uHδy (t)≤ uHδy (

τit),

so A′′2 ≤ A′′3 ≤ A′′4 where

A′′4 =∫ 1

t=0uHδy (

τit)

dG(t) =∫

∆i

t=0uHδy (

τit)

dG(t)+∫ 1

t=∆i

uHδy (τ

it)

dG(t) .

But τ it+∆i

= τ it +∆i. Hence, letting t ′ = t−∆i and renaming t ′ to t,

∫ 1

t=∆i

uHδy (τ

it)

dG(t) =∫ 1−∆i

t=0uHδy (

τit +∆i

)dG(t +∆i) .

Now let i∗ also be large enough that ∆i∗ <ε

2(1−δ )y[2k3+k4]. Then by Lipschitz-G and since,

by (23), uHδy ∈ [0,(1−δ )y], for all i > i∗,

∣∣A′′2−A′′3∣∣≤ ∣∣A′′2−A′′4

∣∣=

∣∣∣∣∫ 1

t=0uHδy (

τit +∆i

)dG(t)−

∫∆i

t=0uHδy (

τit)

dG(t)−∫ 1−∆i

t=0uHδy (

τit +∆i

)dG(t +∆i)

∣∣∣∣≤∫

∆i

t=0uHδy (

τit)

dG(t)+∫ 1

t=1−∆i

uHδy (τ

it +∆i

)dG(t)

+∫ 1−∆i

t=0uHδy (

τit +∆i

)|ψ (t)−ψ (t +∆i)|dt

≤ (1−δ )yk3∆i +(1−δ )yk3∆i +(1−δ )yk4∆i <ε

2.

For all i > i∗, G in G , H in H , and y ∈ (0,y],∣∣∣EuGHδyi−EuGHδy

∣∣∣< ε as claimed.

We now show that E i [yZ] converges uniformly to E∞ [yZ]. Combined with the preceding

result, this will show that EΠGHδyi converges uniformly to EΠGHδy. First,

E i [yZ] =1/∆i

∑c=1

E i [yZ|t = c∆i] [G(c∆i)−G((c−1)∆i)] =∫ 1

t=0E i [yZ|τ i

t +∆i]

dG(t)

so by the triangle inequality,∣∣E i [yZ]−E∞ [yZ]

∣∣≤ A′′′1 +A′′′2 where

A′′′1 =∫ 1

t=0

∣∣E i [yZ|τ it +∆i

]−E∞

[yZ|τ i

t +∆i]∣∣dG(t)

52

Page 53:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

and A′′′2 =∫ 1

t=0

∣∣E∞[yZ|τ i

t +∆i]−E∞ [yZ|t]

∣∣dG(t). By (90), A′′′1 ≤ y∆′i. Integrating by

parts, E∞ [Z|t] = 1−∫ 1

z=0 H (z|t)dz. Thus, by L-H,∣∣E∞[yZ|t = τ

it +∆i

]−E∞ [yZ|t]

∣∣≤ y∫ 1

z=0

∣∣H (z|τ it +∆i

)−H (z|t)

∣∣dz

≤ yk1

∫ 1

z=0

∣∣τ it +∆i− t

∣∣dz≤ yk1∆i,

so A′′′2 ≤ yk1∆i. Hence, E i [yZ] converges uniformly to E∞ [yZ] as claimed.

As for homogeneity in y, using (26), equation (27) for y = 1 can be rewritten as

dDHδ1

dt=

11−δ

∫ DHδ1

z=0∂H(z|t)

∂ t dz

1−H(DHδ1|t

) .Hence, for any solution DHδy to CPHδy, DHδ1 = y−1DHδy is a solution to CPHδ1. But both

solutions DHδy and DHδ1 are unique as shown above. Hence, DHδy must equal yDHδ1.

Thus, the expected payout pHδy (t) = vHy(

DHδy (t) , t)

must equal vHy(

yDHδ1 (t) , t)

.

But by (24), for any D, vHy (D, t) equals yvH1(

Dy , t)

, so pHδy (t) = ypHδ1 (t), uHδy (t) =

yuHδ1 (t), and EuGHδy = yEuGHδ1 as claimed. These properties hold for ΠHδy and EΠGHδy

as well since E∞ [yZ|t] is homogeneous of degree 1 in y.

As for model i, by (22),

vHyi (D, t) = yvH1i(

Dy, t), (91)

so if DHδ1i solves DPHδ1i, then y−1DHδ1i solves DPHδyi. But both solutions DHδyi and

DHδ1i are unique as shown above. Hence, DHδyi must equal yDHδ1i. Thus, by (91), the

expected payout pHδyi (t) = vHyi(

DHδyi (t) , t)

must equal yvH1i(

DHδ1i (t) , t)= ypHδ1i (t)

and so uHδyi (t) equals yuHδ1i (t) as claimed. These properties hold for ΠHδyi and EΠGHδyi

as well since E i [yZ|t] is homogeneous of degree 1 in y. This completes the proof of

Theorem 1. Q.E.D.Theorem 1

4 Relaxing Monotonicity

We now show that the monotonicity assumed in ASSUMPTION A is not necessary for our

results, in the case of two assets and two types with generic parameters. Why? By

53

Page 54:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

genericity, one type must expect her portfolio to have a higher total payout than the other.

Swapping indices if needed, we can assume this is type 2. Also by genericity, an increase

in the seller’s type from 1 to 2 must raise the expected value of one asset proportionally

more than that of the other. Again swapping indices if needed, we can assume that asset 2

has this property. With respect to this labeling of types and assets, Increasing Informational

Sensitivity (IIS) holds: an increase in the seller’s type from 1 to 2 raises the expected value

of asset 2 proportionally more than that of asset 1. We then show that IIS can be used as an

alternative to monotonicity in establishing a unique intuitive equilibrium of the 2x2 model.

In this equilibrium, type 1 sells her entire portfolio, while type 2 retains first asset 2 and

then, if needed, asset 1, until type 1 is just willing not to imitate her. Hence, as in our base

model with the IIS assumption, the Pecking Order property holds: the asset whose value

rises faster in the seller’s type is retained first. However, there is one contrast. Without

monotonicity, asset 1 may be worth less to type 2 than to type 1.6 If so, the seller uses

only asset 2 to signal her type: both types sell asset 1 in its entirety. Intuitively, type 1

gains more from retaining asset 1 than type 2 does. Hence, if type 2 retains some of asset

1, then she must retain even more of asset 2 in order to credibly signal her type. And such

inefficient signalling is ruled out by the Intuitive Criterion, which we assume.

To take a concrete example, suppose a firm has both bonds and common stock to sell,

and these are secured by a common underlying cash flow (such as the value of its real

assets). Suppose when the firm’s type is high, both the mean and the variance of its cash

flow are higher. The latter effect raises the risk of bond default. If the default effect is

strong enough, the high type can place a lower value on the bonds than the low type. In

this case, the model predicts that the firm will always sell its bonds in their entirety: it will

signal optimism by retaining stock alone.

6As shown below in Claim 9, this is not true of asset 2: it is worth more to type 2 than to type 1, as in the

base model.

54

Page 55:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

4.1 The Model

Let A2NM denote the generic 2x2 asset game without monotonicity, which is as follows.

There are two assets i = 1,2 and the seller has two possible types t = 1,2.7 Each type

has positive prior probability. The seller owns a single unit of each asset: the endowment

vector a is (1,1). Let

fi (t)> 0 (92)

denote the expected payout of asset i = 1,2 conditional on the seller’s type being t = 1,2.

We restrict to generic parameters so we can assume, w.l.o.g., that type 2 has a higher

conditional expected portfolio value than type 1:

f1 (2)+ f2 (2)> f1 (1)+ f2 (1)> 0. (93)

(If not, swap type indices.) Genericity also lets us assume, w.l.o.g., that

f2 (2)/ f2 (1)> f1 (2)/ f1 (1)> 0. (94)

(If not, swap asset indices.) Intuitively, (94) says that an increase in the seller’s type raises

the expected value of asset 2 proportionally more than that of asset 1.

Under our assumptions, asset 1 may be worth less to type 2 than it is to type 1: f1 (2)

may be less than f1 (1). On the other hand, asset 2 must be worth more to type 2 than it is

to type 1:

Claim 9. Assume (93) and (94). Then

f2 (2)> f2 (1) . (95)

Proof. If instead f2 (1) ≥ f2 (2) then f2 (2)/ f2 (1) ≤ 1 whence, by (94), f1 (2)/ f1 (1) < 1

so f1 (1)> f1 (2). Combining the first and last inequalities yields f1 (1)+ f2 (1)> f1 (2)+

f2 (2), which contradicts (93).

7We let the type numbers start at one rather than than at zero so as to mirror the asset numbering.

55

Page 56:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

In all other respects, A2NM is simply the Asset Sale game specialized to the 2x2 case.

On seeing her type t, the seller chooses quantities q = (q1,q2) ∈ [0,1]2 to offer for sale as

well as price caps p = (p1, p2). A competitive set of deep-pocketed, uninformed investors

see this choice (q, p) and form posterior beliefs µ (t|q, p) about the probability of each type

t. The seller then sells qi units of each asset i = 1,2 to the investors for some price pi ≤ pi;

let p= (p1, p2) denote the price vector. The payoff of a seller of type t is then q [p−δ f (t)]:

her issuance revenue pq less the present discounted value δq f (t) of the assets she sells.

This completes the specification of A2NM.

In A2NM, the Recursive Linear Program (RLP) takes the following simple form. For

any x ∈ [0,1], define

∆x = (∆x

1,∆x2) = f (2)− x f (1) . (96)

We will often write ∆ and ∆i as shorthand for ∆δ and ∆δi ,respectively:

∆ = (∆1,∆2) = f (2)−δ f (1) where ∆i = fi (2)−δ fi (1) . (97)

By (95), ∆2 is positive; however, ∆1 may be of either sign.

RLP Let Q1 equal the set Q = [0,1]2 of all quantity vectors and let

Q2 = q : q∆≤ u∗ (1) (98)

be the set of all such vectors that can be assigned to type 2 which, if accompanied by

revenue ρ = q f (2), do not tempt type 1 to imitate type 2. For t = 1,2, let

q∗ (t) ∈ argmaxq∈Qt

[q f (t)] , (99)

ρ∗ (t) = q∗ (t) f (t) , and (100)

u∗ (t) = ρ∗ (t)−δq∗ (t) f (t) = (1−δ )q∗ (t) f (t) . (101)

Intuitively, ρ∗ (t) and u∗ (t) are the revenue and payoff functions of a type-t seller from the

quantity vector q∗ (t) under fair pricing: when the resulting price vector is f (t). We now

characterize the solution to RLP with a sequence of results.

56

Page 57:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Claim 10. Any solution to RLP has the following properties.

1. Type 1 sells his whole portfolio:

q∗ (1) = a = (1,1) , (102)

yielding revenue ρ∗ (1)= f1 (1)+ f2 (1) and seller’s payoff u∗ (1)= (1−δ ) [ f1 (1)+ f2 (1)].

2. Type 1’s incentive compatibility (IC) constraint binds at type 2’s quantity vector

q∗ (2):

q∗ (2)∆ = u∗ (1) . (103)

3. Type 2 sells all of asset 1 if she sells any of asset 2: if q∗2 (2)> 0, then q∗1 (2) = 1.

4. A type 2 seller retains a portion of asset 2: q∗2 (2)< 1.

5. The seller’s revenue and payoff are decreasing in her type: ρ∗ (1) > ρ∗ (2) > 0 and

u∗ (1)> u∗ (2)> 0.

The intuition for part 2 is simple: if type 1’s IC constraint does not bind, then type 2

can sell a bit more of either asset without being mistaken for type 1, so she will do so. Parts

2 and 3 also yield an algorithm for finding type 2’s optimal quantity vector. We start with

q= a and gradually lower q2 until it reaches zero, and then begin lowering q1. By part 2, we

must stop when type 1 is just willing not to imitate and be mistaken for 2: at which equation

(103) holds. Hence, a seller does not sell any of asset 2 until she has liquidated all of asset

1. Thus, the Pecking Order property carries over to the 2x2 case without monotonicity. The

algorithm implied by parts 2 and 3 pins down the quantities q∗ (2) that type 2 sells:

Claim 11. In any solution to RLP, the quantities q∗ (2) sold by type 2 are unique and given

by

q∗1 (2) = min

1,u∗ (1)

∆1

∈ (0,1] (104)

and

q∗2 (2) = max

0,u∗ (1)−∆1

∆2

∈ [0,1) . (105)

57

Page 58:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Claim 11 has an interesting implication. If the high-type seller does not value asset 1

more than the low type, then both types sell asset 1 in its entirety. This is the qualitative

impact of discarding ASSUMPTION A. In particular, type 2 retains only asset 2 in order

to signal her information. Intuitively, if type 2 does not value asset 1 more highly than type

1, then retaining it does not help her separate from type 1. As retention is costly, she will

thus sell asset 1 in its entirety.

Corollary 1. If the expected value of asset 1 is not increasing in the seller’s type, then

neither type retains any shares of this asset.

Proof. By (102), type 1 sells asset 1 in its entirety. As for type 2,

u∗ (1)−∆1 = (1−δ ) [ f1 (1)+ f2 (1)]− f1 (2)+δ f1 (1)

= f1 (1)− f1 (2)+(1−δ ) f2 (1) .

If f1 (2)≤ f1 (1), the second line is positive and thus q∗1 (2) = 1 by (104).

The claim can also be proved without relying on (104). If type 2 retains part of asset 1 in

equilibrium (if q∗1 (2)< 1) then she must also retain asset 2 in its entirety (q∗2 (2)= 0) by part

3 of Claim 10. Now suppose type 1 imitates type 2: he retains all of asset 2 as well as part

of asset 1, without obtaining a higher price for asset 1 than he gets in equilibrium (since

f1 (2) ≤ f1 (1)). His payoff from this deviation must be less than (1−δ ) f1 (1) which,

in turn, is less than his equilibrium payoff of u∗ (1) = (1−δ ) [ f1 (1)+ f2 (1)]. Since the

deviation is strictly worse for type 1, his IC constraint does not bind, which contradicts part

1 of Claim 10.

The preceding results have the following useful implication.

Corollary 2. In A2NM, RLP has a unique solution (q∗,ρ∗,u∗).

Proof. By part 1 of Claim 10 and by Claim 11, the quantity function q∗ is unique. Using

(100) and (101), this then pins down the princial’s revenue and payoff functions, ρ∗ and

u∗.

58

Page 59:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

We now prove that the unique solution to RLP is also the unique intuitive outcome. A

perfect Bayesian equilibrium (PBE) of A2NM is a profile

e = (q(·) , p(·) ,µ (·|·, ·) , p(·, ·))

of strategies and beliefs that satisfies the following properties.8

Profit Maximization. The seller’s behavior is optimal given the investors’ price function:

for each t = 1,2, (q(t) , p(t)) ∈ argmaxq,p (q [p(q, p)−δ f (t)]).

Competitive Pricing. Given the investors’ beliefs, an asset’s price equals its conditional

expected value or the price cap, whichever is lower:

p(q, p) = p∧

[2

∑t=1

f (t)µ (t|q, p)

]. (106)

Rational Updating. Investors’ beliefs µ (·|q, p) are given by Bayes’s Rule if some type

chooses (q, p) in equilibrium.

Clearly, any PBE e also induces a reduced-form price function p(t)= p(q(t) , p(t)), revenue

function ρ (t) = q(t) p(t) , and payoff function u(t) = ρ (t)−δq(t) f (t) of the seller.9

In this setting, the Intuitive Criterion is as follows. Fix a PBE e of A2NM with seller’s

payoff function u(·). Suppose a seller of type t deviates to (q, p). The deviation must harm

her if her equilibrium payoff u(t) exceeds her maximum payoff q [p−δ f (t)] from the

deviation - or, equivalently, if her opportunity cost u(t)+δq f (t) of deviating exceeds her

maximum deviation revenue qp. The Intuitive Criterion states that on seeing a deviation

(q, p), investors put zero weight on any type t who is definitely harmed by the deviation if

there is some other type s who is not definitely harmed:

8The notation V ∧V ′ and V ∨V ′ denotes componentwise minimum and maximum, respectively, for any

vectors V and V ′ of equal length.

9We ignore the investors’ payoff function q(t) f (t)− ρ (t), which is identically zero in any intuitive

equilibrium.

59

Page 60:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

The Intuitive Criterion. A PBE e with seller’s payoff function u(·) is intuitive if, for any

deviation (q, p) and any type t such that u(t) + δq f (t) > pq, investors’ posterior

weight µ (t|q, p) on type t is zero if there exists a type s 6= t for whom u(s) +

δq f (s)≤ pq.

We now embed the (by Corollary 2) unique quantity function q∗ in a profile

e∗ = (q∗ (·) , p∗ (·) ,µ∗ (·|·, ·) , p∗ (·, ·)) ,

which we will show is an intuitive PBE; moreover, the outcome of any intuitive PBE is the

same as that of e∗. Other than q∗ (·), the components of e∗ are as follows.

1. Any price caps may be chosen on the equilibrium path as long as they do not bind

for any possible beliefs: p∗ (t)≥ f (1)∨ f (2) for each type t.

2. The price function p∗ (·, ·) is given by (106) with investors’ beliefs µ∗ replacing µ .

Hence Competitive Pricing holds.

3. Investors’ beliefs µ∗ are given by Bayes’s Rule in equilibrium: investors believe that

the seller is of type 1 (resp., 2) for sure if they see the issuance choice (q∗ (1) , p∗ (1))

(resp., (q∗ (2) , p∗ (2))). Thus, Rational Updating holds. Since, moreover, the price

caps do not bind, each equilibrium price p∗i (q∗ (t) , p∗ (t)) equals the conditional

expected payout fi (t) of asset i: the profile e∗ is fairly priced. It follows from

(100) and (101) that the type-t seller’s revenue and payoff functions are ρ∗ (t) and

the payoff u∗ (t) given by RLP. If the seller chooses a pair (q, p) that never occurs

in equilibrium, investors’ beliefs µ∗ (·|q, p) put positive weight only on types t that

have the lowest opportunity cost of giving up their equilibrium payoff u∗: those in

the set T (q) = argmint ′ [u∗ (t ′)+δq f (t ′)]. Hence, beliefs are intuitive. In particular,

if the set T (q) is a singleton t, then investors believe the seller is of type t. If T (q)

contains both types t = 1,2, investors believe the seller is of that type t for which

the deviation revenue q [p∧ f (t)] under symmetric information is minimized. If this

deviation revenue is the same for the two types, then investors’ beliefs do not matter.

60

Page 61:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

We can assume, e.g., that they believe that t = 1, although any other beliefs would

give the same results.

The above specification and argument implies that, in e∗, a type-t seller sells the quantities

q∗ (t) and gets the revenue ρ∗ (t) and payoff u∗ (t) that solve RLP. It remains only to show

that e∗ is the unique intuitive PBE. More precisely, (a) it is an intuitive PBE, and (b) any

other intuitive PBE has identical equilibrium behavior and payoffs to e∗.10

Proposition 1. The profile e∗ is an intuitive PBE.

Proposition 2. If e=(q(·) , p(·) ,µ (·|·, ·) , p(·, ·)) is an intuitive PBE of A2NM, with outcome

u(·) and price function p(·), then the seller makes the same quantity choices and receives

the same payoffs in e as in e∗: for each type t, q(t) = q∗ (t) and u(t) = u∗ (t). Moreover,

if the quantity qi (t) = q∗i (t) of an asset i sold by a type t is positive, the resulting price

of asset i is the same in the two equilibria and equals the fair value of the asset: pi (t) =

p∗i (t) = fi (t).

4.2 Relaxing Monotonicity: Proofs

PROOF OF CLAIM 10. Part 1. First, a f (2) > a f (1) by (93) whence a /∈ Q2 by (98).

Hence either q∗1 (2) or q∗2 (2) is less than one; assume w.l.o.g. that q∗2 (2)< 1. Thus, if (103)

does not hold then for ε in (0,1−q∗2 (2)], the vector q′ = q∗ (2)+ (0,ε) is also in Q2. By

(98), q′ offers type 2 a higher payoff under symmetric information than q∗ (2) does. Hence

q∗ (2) is not optimal for type 2 in Q2, a contradiction.

Part 2. Suppose not: q∗1 (2) < 1 and q∗2 (2) > 0. We will derive a contradiction. Let

ε > 0 be small enough that q∗1 (2)+ ε < 1 and q∗2 (2)− ι > 0 where ι = ε∆1/∆2, whose

denominator ∆2 is positive as noted above. Consider the alternative quantity vector q =

10The other PBE may differ in ways that are not payoff relevant. In particular, it may have different

nonbinding price caps and different beliefs following deviations (as long as these beliefs are intuitive and

serve to deter the deviations). Moreover, if a seller does not sell any shares of a given asset, then the price of

this asset is indeterminate and hence may vary across equilibria.

61

Page 62:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

(q1,q2) where q1 = q∗1 (2)+ ε and q2 equals the lesser of q∗2 (2)− ι and one. First, u∗ (1)−

q∆ can be written as the sum of u∗ (1)− q∗ (2)∆, which is nonnegative since q∗ (2) is

in Q2, and [q∗ (2)−q]∆, which by (95) is bounded below by −ε∆1 + ι∆2 which equals

zero by definition of ι . Hence u∗ (1) ≥ q∆, whence q is in Q2. Moreover, the effect on

type 2’s payoff from switching from q∗ (2) to q (under symmetric information) equals the

difference 1−δ in discount factors times the change in revenue q f (2)−q∗ (2) f (2) which,

in turn, equals minε f1 (2)− ι f2 (2) ,ε f1 (2)+1−q∗2 (2). But both elements in the min

are positive: the first by (94)11 and the second by (92) and since ε > 0 and q∗2 (2)≤ 1. We

conclude that q, which is in Q2, is better for 2 than q∗ (2) under symmetric information,

which contradicts the definition of q∗ (2).

Part 3. This holds by part 2 and since, as shown in the proof of part 1, q∗ (2) 6= a.

Part 4. Together with (92), part 3 implies δq∗ (1) f (1) = δa f (1)> δq∗ (2) f (1) which,

substituted into (103), implies q∗ (1) f (1)> q∗ (2) f (2) as claimed. Multiplying each side

by 1− δ , we obtain u∗ (1) > u∗ (2). Clearly, u∗ (2) ≥ 0 since a type 2 seller can always

set q = (0,0). Finally, let ε = u∗(1)a∆

. The denominator is positive by (93). The numerator

is positive by (101) and (102) and less than the denominator by these equations together

with (93). Hence, ε ∈ (0,1) whence the quantity vector (ε,ε) is feasible. Moreover, this

vector is in Q2 since (ε,ε)∆ = εa∆ = u∗ (1) and by (98). It follows that u∗ (2) is not less

than (1−δ )(ε,ε) f (2), which is positive by (92). Hence u∗ (2) and u∗(2)1−δ

= q∗ (2) f (2) are

positive as claimed. Q.E.D.Claim 10

PROOF OF CLAIM 11. Let qi = q∗i (2) for assets i = 1,2. By part 1 of Claim 10, type 1’s

IC constraint binds so

u∗ (1) = q1∆1 +q2∆2. (107)

11By (94), δ f1 (2) f2 (1)< δ f2 (2) f1 (1), whence

f1 (2)∆2 > f2 (2)∆1

which can be rewritten as f1 (2)∆δ2 > f2 (2)∆δ

1 . This, in turn, implies ε f1 (2) > ε f2 (2)∆δ1/∆δ

2 = ι f2 (2) as

claimed.

62

Page 63:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

We know from part 3 of Claim 10 that q2 < 1. There are now three cases.

1. Say q1 < 1. Then q2 = 0 by part 3 of Claim 10. Thus, by (107), u∗ (1) = q1∆1 whence

q1 6= 0 and thus q1 > 0 which implies ∆1 > 0. For q1 < 1 we require u∗ (1) < ∆1

(which implies ∆1 > 0).

2. Say q2 > 0. Then q1 = 1 by part 3 of Claim 10. Thus, by (107), q2∆2 = u∗ (1)−∆1

so, since ∆2 > 0 by (95), u∗ (1)> ∆1.

3. Say that neither #1 nor #2 holds: q1 = 1 and q2 = 0. Then by (107), u∗ (1) = ∆1.

As the above three conditions u∗ (1) T ∆1 are mutually exclusive and exhaustive, we can

replace ”if” in cases 1-3 by ”if and only if”. This implies the following three mutually

exclusive possibilities, which corresponds respectively to cases 1-3 above.

1. If ∆1 > u∗ (1) then q∗1 (2) = u∗ (1)/∆1 ∈ (0,1) and q∗2 (2) = 0.

2. If ∆1 < u∗ (1) then q∗1 (2) = 1 and q∗2 (2) =u∗(1)−∆1

∆2∈ (0,1).

3. If ∆1 = u∗ (1) then q∗1 (2) = 1 and q∗2 (2) = 0.

This result is stated more succinctly in equations (104) and (105). Q.E.D.Claim 11

PROOF OF PROPOSITION 1. Beliefs are intuitive by construction and we verify Competitive

Pricing and Rational Updating in the text. To show that e∗ is an intuitive PBE, it thus

remains only to verify Payoff Maximization: that neither type has a (strictly) profitable

deviation. By construction, investors’ beliefs µ∗ following a deviation (q, p) depend on q

but not on p. Hence, if a deviator chooses binding price caps, this can only lower his payoff

from deviating. We can thus assume, w.l.o.g., that a deviator chooses price caps that can

never bind (such as the caps p∗).

We begin with type 1. Let

Γ(q) = q∆−u∗ (1) (108)

63

Page 64:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

denote the change in type 1’s payoff from deviating to q and being mistaken for type 2. Let

Λ(q) = [u∗ (2)+δq f (2)]− [u∗ (1)+δq f (1)] (109)

denote the difference between type 2’s and type 1’s opportunity cost of deviating to q. By

construction of µ∗, investors must believe the seller is of type 1 (resp., 2) when Λ(q) is

positive (negative). And when Λ(q) = 0, they think he is of type 1: beliefs are governed by

deviation revenue q f (t) which by (109) and part 3 of Claim 10, is lower for type 1 in this

case. By (97), (101), and (103),

Γ(q∗ (2)) = Λ(q∗ (2)) = 0. (110)

Now suppose type 1 deviates to (q, p). If Λ(q)≥ 0, investors will think the seller’s type

is 1: he gets (1−δ )q f (1) which cannot exceed his equilibrium payoff u∗ (1) as q≤ a. So

such a deviation is not profitable. If instead Λ(q) < 0, investors will think his type is 2

whence his payoff changes by Γ(q), which is nonpositive by the following result.

Claim 12. For any q in [0,1]2, if Λ(q)< 0 then Γ(q)≤ 0.

PROOF OF CLAIM 12. Suppose not: Λ(q)< 0 and Γ(q)> 0. We will derive a contradiction.

For brevity let q denote q∗ (2). Recall that ∆xi is defined in (96). By (92) and (95),

∆x2 > 0 (111)

for all such x, whence

ddx

∆x1

∆x2

∝ f1 (2) f2 (1)− f2 (2) f1 (1)< 0 (112)

by (94). We will rely on the following lemma.

Lemma 5. Let q′ ∈ [0,1]2 be such that Γ(q′)≥ 0 and Λ(q′)< 0. Then q′1 > q1.

Proof. By (110), Γ(q) = 0≤ Γ(q′) whence by (108),

(q1−q′1

) ∆δ1

∆δ2+(q2−q′2

)≤ 0. (113)

64

Page 65:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Also by (110), Λ(q) = 0 > Λ(q′) whence, by (109),(q1−q′1

) ∆11

∆12+(q2−q′2

)> 0. (114)

Subtracting (113) from (114) yields (q1−q′1)(

∆11

∆12− ∆δ

1∆δ

2

)> 0, which by (112) yields q′1 >

q1.

Since Γ(q)> 0, q∆δ > u∗ (1)> 0 by (108) and part 4 of Claim 10. Let α = u∗ (1)/[q∆δ

]∈

(0,1) and q = αq. Since Λ(q)< 0 by assumption, u∗ (1)−u∗ (2) (which is positive by part

4 of Claim 10) exceeds δq∆1 and thus also exceeds δ q∆1 since q = αq and α ∈ (0,1).

Rearranging, we obtain Λ(q)< 0. Since, moreover, Γ(q) = 0 by (108) and (96), Lemma 5

implies that q1 > q1. Hence, by (112),

(q1− q1)

(∆0

1

∆02−

∆δ1

∆δ2

)> 0. (115)

We can also rearrange Γ(q) = Γ(q) = 0 to obtain

(q1− q1)∆δ

1

∆δ2+ q2− q2 = 0. (116)

Adding (115) and (116) yields (q1− q1)∆0

1∆0

2+ q2− q2 > 0. This can be rearranged (with the

result multiplied by 1− δ ) to yield (1−δ ) q∆0 > (1−δ ) q∆0. Moreover, q is in Q2 since

Γ(q) = 0. This contradicts the definition of q as the vector in Q2 that maximizes type 2’s

symmetric-information payoff. Q.E.D.Claim 12

We have shown that a type 1 seller will not deviate. Let us now consider a deviation

(q, p) by a type 2 seller. There are two cases.

1. Λ(q) ≥ 0. Then on seeing q, investors believe the seller’s type is 1. Hence, the

deviation is profitable for the type 2 seller only if

0 < Ω(q) = q [ f (1)−δ f (2)]−u∗ (2) .

If this holds then, by (109), 0 < Λ(q)+Ω(q) = (1−δ )q f (1)−u∗ (1) which is not

possible by (92) and since q≤ a: the deviation is not profitable.

65

Page 66:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

2. Λ(q)< 0. Then on seeing q, investors believe the seller’s type is 2. Hence, type 2’s

change in payoffs from this deviation equals Γ(q), which is nonpositive by Claim

12: the deviation is not profitable.

As this verifies Payoff Maximization, e∗ is an intuitive PBE. Q.E.D.Proposition 1

PROOF OF PROPOSITION 2. We first show that in any intuitive PBE, a seller of a given

type is paid the expected value (given her type) of the portfolio that she sells. The precise

property is as follows.

Fair Pricing. A PBE e with quantity function q(·) and price function p(·) is Fairly Priced

if, for each type t, p(t) f (t) = q(t) f (t).

Lemma 6. Any intuitive PBE e is Fairly Priced.

Proof. Let e be intuitive, and suppose type t chooses (q, p) in equilibrium and that investors

respond with price vector p. First suppose the issuance is underpriced: pq < q f (t). Let S

denote the set of types s for which q f (s)< q f (t). There are two cases.

1. S is empty. Then by Competitive Pricing, if the type t seller deviates to (q, f (t)), her

revenue cannot be less than q f (t). As this would be a profitable deviation, this case

is not possible.

2. S is nonempty. Let s′ ∈ S be a type s for which q f (s) is at a maximum among all s

in S: there is no type s in S for which q f (s) exceeds q f (s′). Then q f (s′) < q f (t)

whence, since pq < q f (t), there is a λ ∈ [0,1) such that

q [p−δ f (t)]q [ f (t)−δ f (t)]

< λ <q [p−δ f (s)]

q [ f (t)−δ f (s)]. (117)

Let type t deviate to (λq, f (t)). By (117), λq [ f (t)−δ f (t)] exceeds q [p−δ f (t)],

which equals u(t). And for each s in S, λq [ f (t)−δ f (s)] is less than q [p−δ f (s)]

which, in turn, is not greater than u(s). Accordingly, for each s in S,

u(t)+δ (λq) f (t)< λq f (t)< u(s)+δ (λq) f (s)

66

Page 67:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

and hence, since s is intuitive, µ (s|(λq, f (t))) is zero for each s in S. Hence, by

Competitive Pricing, the revenue that results from the deviation (λq, f (t)) is λq f (t)

whence type t’s payoff from the deviation is λq [ f (t)−δ f (t)] which, as noted,

exceeds u(t): the deviation is profitable, which is not possible.

We have shown that in an intuitive PBE e, underpricing cannot occur (except perhaps off the

equilibrium path). Hence, pq≥ q f (t) for any type t who chooses (q, p) in equilibrium. But

by Competitive Pricing, pq = q(

p∧[∑

2t=1 f (t)µ (t|q, p)

])≤ ∑

2t=1 q f (t)µ (t|q, p) . Thus,

pq is bounded above by a convex combination of q f (t) for those types t who choose (q, p)

in equilibrium. But by the prior result, pq is also bounded below by q f (t) for each such

type t, so it cannot be less than the highest such q f (t). Hence all types who choose (q, p)

in equilibrium have the same value of q f (t), and this also equals the revenue pq that they

receive: the equilibrium is Fairly Priced.

Moreover, if a PBE is Fairly Priced, then each asset that is sold in equilibrium must be

priced correctly.

Claim 13. Suppose a PBE e with quantity function q(·) and price function p(·) is Fairly

Priced. Then for each type t, and each asset i for which qi (t)> 0, we have pi (t) = fi (t).

Proof. Suppose not. Then by Competitive Pricing, there is a type t who sells some shares

of an asset i with a binding price cap: qi (t) > 0 and pi (t) < fi (t). Thus, this type’s

revenue pi (t)qi (t) from asset i is less than the value qi (t) fi (t) of shares sold. And by

Competitive Pricing, her revenue p j (t)q j (t) from any other asset j cannot exceed its

true value q j (t) f j (t). Combining these facts, we find that her total issuance revenue

p(t)q(t) must be less than the total value q(t) f (t) of shares she sells, which violates

Fair Pricing.

Now let e be an intuitive PBE with outcome u(·) and price function p(·). By Lemma

6, e is Fairly Priced, whence

u(t) = (1−δ )q(t) f (t) for each type t. (118)

67

Page 68:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Hence, u(1) ≤ (1−δ )a f (1) = u∗ (1). And as e is a PBE, type 1 cannot prefer to imitate

type 2. So since e is Fairly Priced, type 2’s choice q(2) must satisfy q(2)∆ ≤ u(1).

As u(1) ≤ u∗ (1), this implies that q(2) lies in Q2. But u∗ (2) is type 2’s highest payoff

(1−δ )q f (2) among all vectors q in Q2, whence u(2)≤ u∗ (2) as well.

Now suppose that in e, type 1 deviates to (q, p) where q = a and p = f (1)∨ f (2).12 By

Competitive Pricing and part 4 of Claim 10, his resulting revenue is a[∑

2t=1 f (t)µ (t|q, p)

]≥

a f (1) whence his payoff from the deviation is at least u∗ (1). Thus, for 1 not to deviate,

u(1) must equal u∗ (1). By (92) and (118), this implies q(1) = a = q∗ (1) as claimed.

Now consider type 2. If u(2) < u∗ (2), suppose type 2 deviates to (q, p) where q =

λq∗ (2) for some λ ∈(

u(2)u∗(2) ,1

)and p = f (1)∨ f (2). By the prior result, u(1) = u∗ (1) =

q∗ (2)∆> λq∗ (2)∆ and u(2)< λu∗ (2)= (1−δ )λq∗ (2) f (2). Thus, u(1)+δλq∗ (2) f (1)>

λq∗ (2) f (2)> u(2)+δλq∗ (2) f (2) whence, since e is intuitive, µ (2|q, p) = 1 and hence,

by Competitive Pricing, type 2’s payoff from the deviation is λu∗ (2) > u(2). Thus, for

type 2 not to deviate, her payoff u(2) must equal u∗ (2). And since u(1) = u∗ (1) and by

Fair Pricing, q(2) must lie in Q2 (else type 1 will deviate to q(2)). So q = q(2) maximizes

type 2’s symmetric-information payoff (1−δ )q f (2) in Q2 whence, by part 6 of Claim 10,

q(2) must equal q∗ (2) as claimed.

Finally, as q(t) equals q∗ (t) for each type t, Claim 13 implies that the prices pi (t) and

p∗i (t) each equal fi (t) for each asset i for which qi (t) = q∗i (t) is positive. Q.E.D.Proposition 2

5 The Maskin and Tirole (MT) Approach

5.1 Preview

This section will test to what extent the properties our Asset Sale (AS) game carry over

to the more flexible sales procedure of Maskin and Tirole (1992), henceforth ”MT”. We

find that when MT’s approach permits outcomes other than the RLP outcome, these other

12Recall that ∨ denotes the componentwise maximum.

68

Page 69:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

outcomes are infinite in number, nonintuitive, and somewhat more likely to involve separation

(as in RLP) than pooling.

We begin, in section 5.2, with the general AS game with T types and n assets. Under

MT’s procedure, the RLP solution (which we will also refer to as the “RSW allocation” for

reasons described below) is the unique intuitive equilibrium. However, other nonintuitive

equilibria cannot be ruled out.

In order to determine when and what sort of other equilibria can exist, we turn to the

case of two types and two assets in section 5.3. The RLP outcome is the unique equilibrium

if the high type seller is not too likely. Otherwise, there also exists a continuum of equilibria

other than RLP. However, many of these nontintuitive equilibria also involve separation

and thus have the same general form as the RLP outcome. Indeed, the set of parameters

for which pooling can occur is a proper subset of the set for which there are non-RLP

separating equilibria. Thus, MT’s approach (without adding the intuitive criterion) does not

yield a unique alternative to RLP, and it actually seems to favor separation over pooling.

In section 5.4 we then apply these results to the Ex Post Security Design Game (DFJ

section 3.3-3.4). While there can be multiple equilibria using MT’s approach, they all

possess the qualitative properties of the RLP outcome identified in DFJ section 3.4: under

the Hazard Rate Ordering property, the issuer sells standard debt with a face value that is

nonincreasing in her type. However, in contrast to the RLP outcome, if the high type is

sufficiently likely then there are outcomes in which the low type issuer is paid more than

the fair value of her security. This loosens her IC constraint, which lets the high type set a

higher face value of his security relative to the RLP outcome. Indeed, if his prior probability

is high enough, this face value can equal that of the low type: the two types can sell the

same standard debt security. Again, none of these non-RLP outcomes are intuitive.

Technically, MT’s key results rely on assumptions do not hold in our AS game. Hence

we must supply proofs that do not use these assumptions. More precisely:

1. MT’s Theorem 1 (which appears below as Proposition 3) assumes the RSW allocation

is interim efficient with respect to some positive beliefs. This assumption lets MT

69

Page 70:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

prove Theorem 1 using known properties of sequential equilibria (MT [12, p. 20]).

As we cannot verify the assumption in our setting, we supply a proof that does not

rely on it. We do so by replacing sequential equilibrium with the more general

solution concept of correlated equilibrium (Aumann [2]) - a change that, in our

setting, does not alter the set of equilibrium outcomes.

2. MT’s Theorem 1 relies on their Proposition 5, which assumes two properties that fail

in our setting:

(a) MT [12, p. 5] write ”We will assume that a type i indifference curve is nowhere

tangent to a type j indifference curve (i 6= j).” Our ASSUMPTION A, in contrast,

permits two adjacent types to have the same information about one or more

assets whence their preferences over these assets coincide.

(b) Later MT [12, p. 11] write ”We will also assume that µ ·(µ ·0)

is not on the

boundary of the feasible set.” In our setting, this means that the quantity qi (t)

sold of each asset, by each type t, lies in the interior (0,ai) of the feasible set.

This condition typically fails in our setting as payoffs are linear. For instance,

under IIS, each quantity sold qi (t) is either 0 or ai with the exception of the

hurdle class quantity qc (t).

Hence we prove MT’s Proposition 5 (which appears below as Claim 21) without the

above two assumptions, relying instead on our ASSUMPTION A (monotonicity).

This section is organized as follows. Section 5.2 studies the general AS game with T types

and n assets. Section 5.3 focuses on the 2x2 case. Section 5.4 applies the results of section

5.3 to the Ex Post Security Design game (DFJ, section 3.3). Omitted proofs appear in

section 5.5.

70

Page 71:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

5.2 The General Case

We first summarize the approach of MT and how it will be applied to the setting of our Asset

Sale (AS) model. We generalize MT by permitting any correlated equilibrium (Aumann

[2]) in the last stage of the game, rather than restricting to Nash equilibrium with a public

coordinating device. This permits us to prove a version of MT’s Theorem 1 that does

not require that the RSW allocation is interim efficient with respect to positive beliefs - a

condition that we may not hold in our setting. While in principal the switch to correlated

equilibrium might expand the set of equilibrium outcomes, in practice it does not.13

As applied to our Asset Sale game, MT’s procedure is as follows. The players consist

of a principal and an agent. The principal corresponds to our seller. The agent is best

interpreted as a single risk-neutral investor with deep pockets, who replaces our continuum

of investors. The payoffs of principal and agent depend on three variables q, ρ , and t,

defined as follows.

1. q is a vector of observable and verifiable actions, belonging to a compact, convex set

Q ⊂ℜn. In our AS model, q is the vector (q1, ...,qn) of quantities that the principal

sells to the agent, and

Q = [0,a1]×·· ·× [0,an] (119)

is the set of all feasible quantity vectors.

2. ρ ∈ ℜ is a monetary transfer from the agent to the principal. In the AS model it is

the revenue ρ = pq that the investors pay the seller. To ensure compactness, we will

restrict ρ to lie in [−ρ,ρ] for some arbitrarily large (but finite) constant ρ .

3. t ∈ ϒ = 0, ...,T denotes the principal’s private information or type, with prior

13In a given game, there can be correlated equilibria that are not Nash equilibria - even if the latter include

a public coordinating device (Aumann [2, pp. 70-71]). However, in MT, the principal has enough flexibility

in designing the game that the Nash restriction does not shrink the set of outcomes relative to correlated

equilibrium.

71

Page 72:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

probability distribution Π· = ΠtTt=0 ∈ ∆T (the standard T simplex). This is just

as in the AS model.

MT define an outcome o as an action-transfer pair (q,ρ):14

Definition 1. An outcome o is a pair (q,ρ) ∈ O where

O = Q×ℜ⊂ℜn+1 (120)

is the set of all possible outcomes. Let qo and ρo denote the action and transfer, respectively,

of the outcome o: that is, o = (qo,ρo).

The principal and agent have Neumann-Morgenstern (vNM) utility functions V t (o) and

U t (o), respectively, over outcomes o. As the notation indicates, both payoff functions also

depend on the principal’s type t. In our AS model, these payoff functions are15

V t (o) = ρo−δqo f (t) and U t (o) = qo f (t)−ρ

o. (121)

Let

o0 =(−→

0 ,0)

(122)

denote the reservation outcome in which no assets or cash are transferred and thus, by

(121), each player gets a realized payoff of zero. Further define:

Definition 2. An allocation is a menu o· = otTt=0 ∈ A of outcomes, one for each type of

principal, where

A = OT+1 ⊂ℜ(n+1)(T+1) (123)

is the set of all possible allocations.

14In MT, an outcome is denoted µ . We use o since, in our setting, µ is used to denote beliefs. Moreover,

MT work with random outcomes, which are probability distributions over pairs (q,ρ). However, as our

players are risk-neutral, they care only about the expected action Eq and transfer Eρ . Hence there is no gain

in generality from working with random outcomes: any result concerning an outcome (q,ρ) will hold also

for any random outcome whose expectation is (q,ρ).

15In MT, the variables q, ρ, and t are called y, t, and i, respectively.

72

Page 73:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Definition 3. An allocation o· is incentive-compatible if, for all types s and t, V t (ot) ≥

V t (os).

Definition 4. An allocation o· Pareto-dominates an allocation o· if V t (ot) ≥ V t (ot) for

each type t, where the inequality holds strictly for some t.

Definition 5. Let S be a set of allocations. An allocation o· is Pareto optimal in S if o· ∈ S

and o· is not Pareto-dominated by any other allocation in S.

On seeing his type, the principal proposes a mechanism m: a set of actions for each

player and a map from action pairs to outcomes. The agent may reject the mechanism, in

which case the no-trade outcome o0 is played. Without loss of generality, we incorporate

this option as a special action in the mechanism that always leads to the no-trade outcome:

Definition 6. A mechanism m consists of a finite set SPm of actions for the principal, a finite

set of actions SAm for the agent, and, for each pair of actions

(sP,sA) ∈ SP

m×SAm, an outcome

osP,sA

m ∈ O, such that the agent’s action set SAm contains a special action sA

0 that prevents

trade: if the agent selects sA0 then, for any principal’s action sP ∈ SP

m, the outcome osP,sA0

m

selected by m is the no-trade outcome o0.

Let M denote the set of mechanisms m. Eliminating an explicit “accept/reject” stage

lets us specify MT’s game simply as follows.

Stage 1. The principal privately observes his type t and announces a mechanism m in the

set M.

Stage 2. The principal and agent simultaneously choose actions sP ∈ SPm, sA ∈ SA

m. Their

realized payoffs are V t(

osP,sA

m

)and U t

(osP,sA

m

), respectively.

Following the principal’s choice of a mechanism m, the agent forms interim beliefs Π· (m)

about the principal’s type, where Πt (m) is the probability that she assigns to type t.

Our solution concept for stage 2 is an adaptation of strategic form correlated equilibrium

(SFCE) (Forges [6]) to our setting. Why must SFCE be altered? An SFCE is defined for

73

Page 74:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

a static, one-shot game in which a player’s beliefs about her opponents’ types are given by

an exogenous prior distribution. Hence, if player 1 believes player 2 has probability zero of

being some type t, then type t indeed cannot occur so the SFCE may assign a suboptimal

action to type t. This feature is not suitable in our setting, where the agent’s beliefs are

given by her interim beliefs Π· (m). In particular, if the mechanism m is unexpected, the

agent’s interim beliefs are arbitrary and so may assign zero probability to a type t who has

positive ex-ante probability - even if it was type t who chose m! Hence, a SFCE would let a

type t deviate to some mechanism m and then play suboptimally in the continuation game,

thus violating the principle of sequential rationality.

We fix this by requiring optimality type-by-type for the principal, regardless of the

agent’s beliefs. More precisely, let

Φm = ∆

∣∣∣(SPm)

T+1×SAm

∣∣∣ (124)

be the set of distributions of pure action profiles (where the principal’s action sP· ∈(SP

m)T+1

specifies the action sPt ∈ SP

m to be taken by each type t).

Definition 7 (CEIGΠ·m ). A distribution πm∈Φm is a correlated equilibrium of the incomplete-

information game with mechanism m and interim beliefs Π·, or “CEIGΠ·m ”, if and only if

1. each type of principal is willing to play each action when told to do so:

∑sA∈SA

m

πm

(sP· ,s

A)

V t(

osPt ,s

A

m

)≥ ∑

sA∈SAm

πm

(sP· ,s

A)

V t(

osPt ,s

A

m

)for any t,sP

· , sP· ∈(SP

m)T+1

;

(125)

2. and the agent is willing to play each action when told to do so:

∑sP· ∈(SP

m)T+1

πm

(sP· ,s

A)

uAm,Π·

(sA,sP

·

)≥ ∑

sP· ∈(SP

m)T+1

πm

(sP· ,s

A)

uAm,Π·

(sA,sP

·

)for any sA, sA ∈ SA

m

(126)

where

uAm,Π·

(sP· ,s

A)=

T

∑t=0

ΠtU t(

osPt ,s

A

m

)(127)

74

Page 75:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

is the agent’s expected payoff from the pure strategy profile(sP· ,s

A) under the beliefs

Πt .

Definition 8. A profile Σ of the full game is a triplet(

p··,Π· (·) ,π·

)where pt

m is the

probability that the principal will choose the mechanism m ∈ M given his type t; Πt (m)

is the agent’s posterior probability that the principal’s type is t given that he chose m; and

πm(sP· ,s

A) is the probability, conditional on the principal choosing mechanism m, that the

pure action profile(sP· ,s

A) will be played in stage 2.

Definition 9. An equilibrium of the full game is a profile Σ =(

p··,Π· (·) ,π·

)that satisfies

the following three conditions:

1. for any mechanism m (including those not chosen in equilibrium), πm is a CEIGΠ(m)·

m ;

2. Bayes’s Rule is used whenever possible: for any mechanism m that some type of

principal chooses in equilibrium, the agent’s posterior beliefs are determined by

Bayes’s Rule:

Πt (m) =

Πt ptm

∑Ts=0 Πs ps

m; and (128)

3. for each type t, each mechanism m in the support of pt· is an optimal choice for the

type-t principal given the map π· from mechanisms to correlated equilibria.

5.2.1 Results

For any mechanism m and beliefs Π· ∈ ∆T , let

φm

(Π·)⊂Φm (129)

be the set of associated correlated equilibria of stage 2,16 and let

Um (πm) = ∑sP· ∈(SP

m)T+1

∑sA∈SA

m

πm

(sP· ,s

A)

uAm

(sP· ,s

A)

(130)

16The set Φm of distributions over pure action profiles is defined in (124).

75

Page 76:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

and

V tm (πm) = ∑

sP· ∈(SP

m)T+1

∑sA∈SA

m

πm

(sP· ,s

A)

uPm,t

(sP· ,s

A)

(131)

be the resulting payoffs of the agent and of the type-t principal, respectively, when they

play πm in the mechanism m. Let

ψm

(Π·)=(

V ·m (πm) ,Um (πm))

: πm ∈ φm

(Π·)⊂ℜ

T+1×ℜ (132)

denote the set of payoff vectors(V ·,U

)that can occur in correlated equilibria πm given the

mechanism m and interim beliefs Π·.

Claim 14. For any mechanism m and interim beliefs Π·, the agent’s expected payoff Um (πm)

(evaluated using Π·) in any correlated equilibrium πm ∈ φm

(Π·)

is nonnegative.

Proof. The agent can get zero by always ”rejecting” in stage 2; hence, her expected stage-2

continuation payoff Um (πm) in any correlated equilibrium πm in φm

(Π·)

must be nonnegative.

A simple but important type of mechanism m is a Direct Revelation Mechanism DRM.

Definition 10. A Direct Revelation Mechanism (DRM) is a mechanism m with the following

properties. The agent’s action set SAm consists of two actions: accept or reject. The principal’s

action set SPm is just her type space ϒ. If the agent rejects then, for any action ”t” of

the principal, the reservation outcome o0 is implemented: each player gets a payoff of

zero. If the agent accepts and the principal chooses ”t”, then the DRM implements some

prespecified outcome ot .

For the principal, choosing the action t ∈ ϒ is interpreted as claiming to be of type t. A

DRM m is fully specified as the allocation o· = otTt=0 that is implemented when the agent

accepts and the principal truthfully reports his type. Thus, we will also refer to a DRM

as the allocation o· that it is designed to implement. We will also say that the principal

chooses the outcome ot in stage 2; this means that the principal reports the type ”t”.

A DRM equilibrium o· is an equilibrium in which the allocation o· is implemented as a

DRM. More precisely:

76

Page 77:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Definition 11. A DRM equilibrium o· is an equilibrium in which, as long as there has been

no prior deviation, each type of principal chooses the same DRM o· in stage 1 while, in

stage 2, the agent chooses ”accept” and the principal of each type t truthfully reports his

type: the allocation o· is implemented.

Fix an equilibrium Σ =(

p··,Π· (·) ,π·

). By risk neutrality, Σ gives the principal (resp.,

agent)a type-contingent expected payoff of V t (otΣ

)(resp., U t (ot

Σ

)) when the principal’s

type is t, where17

otΣ = ∑

m∈Mpt

m ∑sP· ∈(SP

m)T+1

∑sA∈SA

m

πm

(sP· ,s

A)

osPt ,s

A

m (133)

is the expected outcome for this type. We will refer to o·Σ

as the expected allocation of Σ.

Definition 12. The allocation o· is an equilibrium allocation if there exists an equilibrium

Σ whose expected allocation o·Σ

is o·.

We next show that for every equilibrium Σ, there is a DRM equilibrium that implements

the expected allocation o·Σ

of Σ. Since a DRM equilibrium is also an equilibrium, this

means that o· is an equilibrium allocation if and only if there is a DRM equilibrium that

implements it. This will help us characterize the set of equilibrium allocations.

Claim 15. For any equilibrium Σ, there is a DRM equilibrium o·Σ.

Proof. Section 5.5.

The following definition is from MT [12, p. 10].

Definition 13. An allocation o· is weakly interim efficient (WIE) if and only if (a) it is

incentive compatible -

V t (ot)≥V t (os) for all types t and s (134)

17For any outcomes o = (q,ρ) and o′ = (q′,ρ ′), and any probability π , the expected outcome πo +

(1−π)o′ refers to the outcome in which the portfolio πq+(1−π)q′ is transferred in return for the payment

πρ +(1−π)ρ ′ for sure. By risk neutrality, this deterministic outcome is payoff-equivalent to a lottery in

which outcome o occurs with probability π and outcome o′ occurs with probability 1−π .

77

Page 78:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

- and there exists no allocation o· that (a) Pareto dominates o·:

V t (ot)≥V t (ot) for all types t, strictly for some t; (135)

(b) is incentive-compatible:

V t (ot)≥V t (os) for all types t and s; (136)

and (c) gives the agent at least as high a payoff as o· does for each type of principal:

U t (ot)≥U t (ot) for all types t. (137)

MT [12, p. 10] use also the following equivalent formulation.

Claim 16. An allocation o· is WIE if and only if it solves the following maximization

problem for some vector of positive weights wtTt=0 ∈ℜ

T+1++ :

Program I. maxo·∑Tt=0 wtV t (ot) subject to (136) and (137).

Proof. Section 5.5.

Recall that there is no trade in the reservation allocation o·0: each player gets zero.

Definition 14. An allocation o·, with payoffs

V t d=V t (ot) for all t, (138)

is an RSW allocation if and only if, for all t, it solves:

Program IIt . V t = maxo·V t (ot) subject to

V s (os)≥V s (or) for all types r,s; (139)

U s (os)≥U s (os0) for all types s. (140)

Claim 17. There exists an RSW allocation o·. Any RSW allocation o· is WIE.

Proof. The proof is the same as in MT [12, p. 11].

78

Page 79:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

For convenience, we now restate the Recursive Linear Program (RLP). Recall that Q,

defined in (119), is the set of all quantity vectors q.

RLP Let Q0 = Q and, for each t = 1, ...,T , let

Qt = q ∈ Q : for all s < t, q [ f (t)−δ f (s)]≤ u∗ (s) (141)

be the set of all quantity vectors that can be assigned to type t which, if accompanied

by revenue ρ = q f (t), do not tempt any lower type to imitate type t. For each

t = 0, ...,T , let

q∗ (t) ∈ argmaxq∈Qt

[q f (t)] , (142)

ρ∗ (t) = q∗ (t) f (t) , and (143)

u∗ (t) = ρ∗ (t)−δq∗ (t) f (t) . (144)

RLP is thus a maximization program that produces a seller’s payoff function u∗, a quantity

function q∗, and a revenue function ρ∗, where u∗ and ρ∗ are unique by DFJ’s PROPOSITION

1.18 Any solution to RLP yields an associated allocation o· where ot = (q∗ (t) ,ρ∗ (t)). In

any such allocation, the payoff q∗ (t) f (t)−ρ∗ (t) of investors is zero by (143).

Crucially, RSW and RLP are equivalent in our setting, in the following sense.

Claim 18. There is a one-to-one correspondence between RSW allocations and solutions

to RLP, which is as follows.

1. Let o· be an RSW allocation. Then the agent’s payoff U t (ot) is identically zero and

the functions q∗ (t) = qot, ρ∗ (t) = ρ ot

, and u∗ (t) =V t (ot) solve RLP.

2. Let (u∗,ρ∗,q∗) solve RLP. Then the allocation o· with ot = (qt ,ρ t) = (q∗ (t) ,ρ∗ (t))

is an RSW allocation.

18Multiple quantity functions q∗ may solve RLP. E.g., if the two assets i and j have the same conditional

expected values (if fi (t) = f j (t) for all t) then the players care only about their total holdings of the two

assets combined. Hence, RLP cannot tell us the number of shares a seller of a given type t will sell of each

asset, although it may pin down the sum qi (t)+q j (t).

79

Page 80:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Proof. Section 5.5.

The following definition is from MT [12, p. 10].

Definition 15. An allocation o· is interim efficient (IE) relative to beliefs Π· if and only if

(a) it is incentive compatible -

V t (ot)≥V t (os) for all types t and s (145)

- and there exists no allocation o· that (a) Pareto dominates o·:

V t (ot)≥V t (ot) for all types t, strictly for some t; (146)

(b) is incentive-compatible:

V t (ot)≥V t (os) for all types t and s; (147)

and (c) gives the agent at least as high expected payoff as o· under the beliefs Π·:

T

∑t=0

ΠtU t (ot)≥ T

∑t=0

ΠtU t (ot) . (148)

MT [12, p. 10] use also the following equivalent formulation.

Claim 19. An allocation o· is IE with respect to the beliefs Π· if and only if it solves the

following maximization problem for some vector of positive weights wtTt=0 ∈ℜ

T+1++ :

Program I. maxo·∑Tt=0 wtV t (ot) subject to (147) and (148).

Proof. Section 5.5.

The following is MT’s Proposition 3, specialized to our model.

Claim 20. For any WIE allocation o·, the set of beliefs Π(o·) for which o· is IE is nonempty

and convex.

Proof. Section 5.5.

80

Page 81:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Theorem 1 in MT relies on the following claim (their Proposition 5). The proof in MT

relies on their assumption that the indifferent curves of different types are never tangent.

In our setting, two adjacent types may have identical preferences. In this case, their

indifference surfaces in (q,ρ) space are identical. Hence, we use a different proof, which

relies on our ASSUMPTION A.

Claim 21. In any equilibrium of the contract proposal game, the payoff of the type t

principal is at least her RSW payoff V t (ot).

Proof. Section 5.5.

We now extend the Intuitive Criterion (Cho and Kreps [4]) to our setting. This criterion

restricts the investor’s beliefs when the principal deviates to a mechanism m that no type’s

strategy puts positive weight on: for all t, ptm = 0. Recall that ϒ is the set of all types t

and V tm (πm), defined in (131), is the type-t principal’s payoff in the correlated equilibrium

πm of the mechanism m. For a fixed equilibrium Σ =(

p··,Π· (·) ,π·

), the payoff V t of a

type-t principal is V t (otΣ

)where ot

Σis the expected outcome for this type, defined in (133).

Consider the following restriction on the agent’s interim beliefs Π· (m) in the equilibrium

Σ, following the principal’s choice of mechanism m, where, for any set S⊂ ϒ of types,

φSm =

⋃Π·∈∆T : supp(Π·)⊂S

φm

(Π·)⊂Φm

denotes the set of correlated equilibria for interim beliefs Π· whose support lies in S.19

IntΣm Fix an equilibrium Σ. For any mechanism m, let Jm ⊂ ϒ denote the (possibly empty)

set of types t for whom the payoff of the principal in Σ exceeds his maximum payoff

in all correlated equilibria of m for any interim beliefs of the agent:

Jm =

t ∈ ϒ : V t (ot

Σ

)> max

πm∈φ ϒm

V tm (πm)

. (149)

19The set φm

(Π·)

of correlated equilibria of m for beliefs Π· is defined at equation (129).

81

Page 82:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Suppose there exists a type t in ϒ− Jm of principal whose payoff in Σ is no higher

than his minimum payoff in any correlated equilibrium of m that results from beliefs

that put zero weight on any type in Jm:

V t (otΣ

)≤ min

πm∈φϒ−Jmm

V tm (πm) . (150)

Then on seeing m, the agent’s interim beliefs Π· (m) put zero weight on any type in

Jm: supp[Π· (m)

]⊂ ϒ− Jm.

IntΣm holds automatically if m is offered by some type in equilibrium. Why? No type in

Jm will choose m, and m is chosen with positive probability in Σ, so by Bayes’s Rule the

agent’s interim beliefs on seeing m must put zero weight on Jm: IntΣm holds.

We use IntΣm to define intuitive beliefs, and an intuitive equilibrium.

Definition 16 (The Intuitive Criterion). Let Σ =(

p··,Π· (·) ,π·

)be an equilibrium. The

beliefs Π· (·) and equilibrium Σ are intuitive if and only if IntΣm holds for every mechanism

m ∈M.

Without strengthening our ASSUMPTION A, we cannot prove the following result,

which is needed to apply MT’s Theorem 1 (stated below as Proposition 3) to our model:

Claim 22. Let o· be an RSW allocation. Then o· is IE with respect to some positive beliefs

Π· ∈ℜT+1++ .20

We will instead rely only on the following weaker property:

Claim 23. Let o· be an RSW allocation. Then o· is IE relative to some nonnegative beliefs

Π· ∈ℜT+1+ .

Proof. Follows from Claims 34 and 17.

We now state and prove our analogue to MT’s Theorem 1. The result shows also that

the RSW allocation is intuitive. Recall that V t , defined in (138), is the RSW payoff of a

principal of type t.

20That is, Πt > 0 for each type t.

82

Page 83:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Proposition 3. An allocation o· is the expected allocation of some equilibrium if and only

if (a) it is incentive compatible:

V t (ot)≥V t (os) for all types t and s; (151)

(b) it is ”profitable in expectation” for the agent under the prior beliefs Π·:

T

∑t=0

ΠtU t (ot)≥ T

∑t=0

ΠtU t (ot

0)

; (152)

and (c) it gives each type of principal at least her RSW payoff:

V t (ot)≥ V t for each type t. (153)

Moreover, if o· is an RSW allocation, there is an intuitive equilibrium with expected allocation

o·.

Proof. Section 5.5.

Proposition 3 shows that any RSW allocation o· is intuitive: it is the expected allocation

of an intuitive equilibrium. The converse also holds:

Proposition 4. Any expected allocation of an intuitive equilibrium is an RSW allocation.

Proof. Section 5.5.

Hence, the sets of intuitive and RSW allocations coincide. Together with Claim 18, this

establishes that the set of intuitive allocations of the above game coincides with the set of

solutions to RLP which, in turn, are the set of intuitive outcomes of our AS game.

We have shown that under the intuitive criterion, the results of our AS game are robust

to the choice of mechanism: it does not matter whether the assets are sold using our

procedure or that of MT. However, without the intuitive criterion, MT’s approach may

permit additional outcomes by Proposition 3.

83

Page 84:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

5.3 The MT Approach: 2x2 Case

We next apply the results of section 5.2 to the case of two assets i = 1,2 and two types

t = 1,2.21 Recall that fi (t) is the expected payout of asset i = 1,2 conditional on the

seller’s type being t = 1,2. In order to apply our prior results on the A2NM model (section

4.1), we assume generic parameters. Hence, by ASSUMPTION A (monotonicity),

fi (2)> fi (1)> 0 for each asset i. (154)

Swapping asset indices if needed, we can also assume w.l.o.g. that

f2 (2)/ f2 (1)> f1 (2)/ f1 (1) . (155)

That is, an increase in the seller’s type raises the expected value of asset 2 proportionally

more than that of asset 1: IIS holds. Finally, we normalize the number of shares of each

asset to one: the seller’s endowment vector a is (1,1).

We first preview our results graphically. Figure 1 illustrates the case in which case the

type 2 seller retains all of asset 2 in the RSW allocation.22 Figure 2 illustrates the case in

which the type 2 seller sells all of asset 1 in the RSW allocation. We now explain these

figures.

In any equilibrium allocation o·, let x denote the difference between the low type’s

equilibrium payoff V 1 (o1) and her RSW payoff V 1. We show that in any equilibrium

allocation other than the RSW allocation, x is positive (Claims 26 and 27): the low type

must improve on her RSW payoff. The low type’s payoff increment x is depicted on the

vertical axis of Figures 1 and 2, while the prior probability Π2 of the high type appears on

the horizontal axis.

Let Sx denote the set of equilibrium allocations that raise the low type’s payoff by

exactly x> 0 vis-a-vis the RSW allocation. For simplicity, we focus on the set Sx of efficient

21We deviate from the type numbering of section 5.2, which starts at zero, in order to more easily reference

the results of section 4.1.

22The notation ν , η , ζ , and γ will be defined later.

84

Page 85:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

0

0.5

1

1.5

2

2.5

3

3.5

4

0.0

0

0.0

2

0.0

5

0.0

7

0.1

0

0.1

2

0.1

4

0.1

7

0.1

9

0.2

2

0.2

4

0.2

6

0.2

9

0.3

1

0.3

4

0.3

6

0.3

8

0.4

1

0.4

3

0.4

6

0.4

8

0.5

0

0.5

3

0.5

5

0.5

8

0.6

0

0.6

2

0.6

5

0.6

7

0.7

0

0.7

2

0.7

4

0.7

7

0.7

9

0.8

2

0.8

4

0.8

6

0.8

9

0.9

1

0.9

4

0.9

6

0.9

8

x:

Lo

w T

yp

e's

Pay

off

In

crem

ent

Π2 : probability of high type

Efficient Equilibria under MT Procedure

2x2 Case, ν > 0

Pooling Equilibria: ! = "

Separating Equilibria with !" = 1, "

" > 0Separating

Equilibria with

!" < 1, "

" = 0

η ζ

ν

γ

No Equilibria in This Region

Figure 1: the case ν > 0.

85

Page 86:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

0

0.5

1

1.5

2

2.5

3

3.5

4

0.0

0

0.0

2

0.0

5

0.0

7

0.1

0

0.1

2

0.1

4

0.1

7

0.1

9

0.2

2

0.2

4

0.2

6

0.2

9

0.3

1

0.3

4

0.3

6

0.3

8

0.4

1

0.4

3

0.4

6

0.4

8

0.5

0

0.5

3

0.5

5

0.5

8

0.6

0

0.6

2

0.6

5

0.6

7

0.7

0

0.7

2

0.7

4

0.7

7

0.7

9

0.8

2

0.8

4

0.8

6

0.8

9

0.9

1

0.9

4

0.9

6

0.9

8

x:

Lo

w T

yp

e's

Pay

off

In

crem

ent

Π2 : probability of high type

Efficient Equilibria under MT Procedure

2x2 Case, ν < 0

Pooling Equilibria: ! = "

Separating Equilibria with !" = 1, "

" > 0

η ζ

γ

No Equilibria in This Region

Figure 2: the case ν ≤ 0.

86

Page 87:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

elements of Sx: the elements in Sx that give the greatest social welfare, which we define as

the sum of the unconditional expected payoffs of seller and agent. If Sx is nonempty, then

Sx is a singleton.23

This unique efficient allocation has two key properties. First, the low type sells his

whole portfolio but gets more than the fair value of this portfolio: ρ1 > a f (1). This has the

effect of loosening his incentive compatibility constraint, so the high type can sell more of

his portfolio. In this way, more gains from trade are realized than in the RSW allocation so

welfare is higher. Second, as in the RSW allocation, the high type sells all of asset 1 before

he sells any of asset 2.24

These properties allow us to divide the efficient allocations (over all possible type-1

payoff increments x) into three categories, all of which appear in Figure 1.25 For a range

of payoff increments x there are pooling allocations, in which the high type also sells his

entire portfolio: q2 = a. For an intermediate range of x, there are separating allocations in

which the high type sells all of asset 1 and some of asset 2 (q21 = 1 and q2

2 > 0). Finally, in

Figure 1 there is a low range of increments x for which there are separating allocations in

which the high type sells some of asset 1 but none of asset 2 (q21 > 0 and q2

2 = 0).26

This third set of allocations is present in Figure 1 but not in Figure 2. Why? As noted,

any efficient allocation yields higher welfare than the RSW allocation. And as also noted,

the high type must sell all of asset 1 before any of asset 2, as in the RSW allocation. It

follows that he cannot sell fewer shares of asset 1 than in the RSW allocation, as he cannot

sell more sales of asset 2 to offset this. Finally, Figure 2 concerns the case in which the

23See Claim 29 and Proposition 5.

24These are, respectively, properties P1 and P4 in Claim 28.

25These results appear in Claim 29 and Proposition 5.

26As shown, each type of allocation exists only if the prior probability Π2 of the high type exceeds a

threshold that is nondecreasing in x. Intuitively, the low type seller’s gain x equals the agent’s loss when she

encounters this type. In order to accept a larger such loss x, the probability 1−Π2 of encountering the low

type must be lower.

87

Page 88:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

high type sells all of asset 1 in the RSW allocation. Hence, in this case, he must also sell

all of asset 1 in any efficient allocation.

These figures show that in the 2x2 model with MT’s sale procedure, pooling is less

likely than separation. More precisely, for any prior probability Π2 of the high type that

permits pooling, there is a continuum of separating equilibria. But in Figure 1, there is also

a positive-measure set of prior probabilities (the interval Π2 ∈ [η ,ζ ]) for which separating

equilibria exist but pooling equilibria do not. These properties are indeed general, as shown

in Claim 29 and Proposition 5 below.

The final result of this section, Claim 30, characterizes the set of globally efficient

equilibrium allocations: those equilibrium allocations that give the highest social welfare

for any payoff increment x. This set consists of the pooling equilibria depicted in Figures

1 and 2, together with the upper, curved boundary of the region of separating equilibria for

Π2 between η and ζ in Figure 1.27

To proceed, let us define the following set

Definition 17. Let S denote the set of allocations o· that are incentive-compatible for the

principal:

V 1 (o1)≥V 1 (o2) and (156)

V 2 (o2)≥V 2 (o1) ; (157)

that give the agent a nonnegative unconditional expected payoff:

U (o·) =2

∑t=1

ΠtU t (ot)≥ 0 (158)

27In stating this result, we do not mean to imply that the globally efficient equilibrium will be selected. The

equilibrium selection literature has not provided any convincing rationale for this “Pareto criterion”; see, e.g.,

the skeptical discussion in Fudenberg and Tirole [7, pp. 20-22]. While we also focus on efficient equilibria

(for a given payoff increment) in Claim 29 and Proposition 5, our purpose there is to not to make predictions

but rather to shrink the equilibrium set to manageable size while still giving a flavor of the wide variety of

outcomes that can occur.

88

Page 89:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

where Πt > 0 is the prior probability of type t; and that give each type of principal at least

her RSW payoff:

V 1 (o1)≥ V 1 and (159)

V 2 (o2)≥ V 2. (160)

Claim 24. 1. S is the set of equilibrium allocations o·.

2. S contains the RSW allocation o·.

3. S is compact.

Proof. Parts 1 and 2 follow from Proposition 3. As for part 3, each qt lies in [0,1]2 whence:

1. by (154), (159), and (160),

ρt ≥ δqt f (t)+V t ≥ V t >−∞ for t = 1,2; (161)

2. by (158), ∑2t=1 Πtρ t ≤ ∑

2t=1 Πta f (t) whence, for s = 1,2,

ρs ≤ (Πs)−1

[−Π

3−sρ

3−s +2

∑t=1

Πta f (t)

]

≤ (Πs)−1

[−Π

3−sV 3−s +2

∑t=1

Πta f (t)

]< ∞ by (161).

Hence, S is compact.

In the next claim, we solve analytically for the RSW allocation and show it is unique.

Under (154), the model we study is a special case of A2NM described in section 4.1. Thus,

the solution to RLP in this section is identical to that characterized by Claims 10 and 11.

And by Claim 18, for each type t, the principal’s RSW payoff Vt equals his RLP payoff

u∗ (t) while the reservation payoff U t (ot0)

of the agent is zero by (121) and (121). This

implies the first part of the following claim. Recall the notation ∆i = fi (2)− δ fi (1) for

i = 1,2 and ∆ = (∆1,∆2) = f (1)− δ f (2) (equation (97)). We will also use the notation

89

Page 90:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

∆1i = fi (2)− fi (1) and ∆1 =

(∆1

1,∆12)= f (1)− f (2) from (96). The following facts will

be useful.

∆12∆1−∆

11∆2 = (1−δ ) [ f1 (1) f2 (2)− f2 (1) f1 (2)]> 0 by (155); (162)

∆i−δ∆1i = (1−δ ) fi (2)> 0 by (154); (163)

Also define28

γ = δ(a− q2)

∆1 ∈(0,a∆

1) (164)

and29

ν = γ−δ∆12 < γ, (165)

as well as

η = δ∆11/∆1 (166)

and

ζ = γ/a∆1. (167)

Claim 25. There is a unique RSW allocation o· =(

ρ1, q1) ,(ρ2, q2), with the following

properties.

q1 = a = (1,1) ; (168)

ρ1 = a f (1) ; (169)

V 1 = (1−δ )a f (1) =V 1 (o2)= q2∆; (170)

q21 = min

1,

V 1

∆1

∈ (0,1] ; (171)

q22 = max

0,

V 1−∆1

∆2

∈ [0,1) ; (172)

ρ2 = q2 f (2) ; (173)

V 2 = (1−δ ) q2 f (2) . (174)

28The second equality holds by (179) and the set inclusion by (154), (171), and (172).

29The inequality holds by (154).

90

Page 91:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

The following properties also hold:

V 2 >V 2 (o1)= a∆; (175)

V 1−V 2 = δ q2∆

1 (176)

= δ

min

1,

V 1

∆1

11 +max

0,

V 1−∆1

∆2

12

; (177)

V 1−V 2 (o1)= δa∆1; (178)

V 2−V 2 (o1)= γ = δ(a− q2)

∆1 = δa∆

1−(

V 1−V 2)

; (179)

ν = δ∆11−(

V 1−V 2)

; (180)

q21 = min

1,

V 1−V 2

δ∆11

; (181)

q22 = max

0,

V 1−V 2−δ∆11

δ∆12

= max

0,−ν

δ∆12

; (182)

V 1 S ∆1 as V 1−V 2 S δ∆11 as ν T 0; (183)

V 1 = a∆−a∆1; (184)

ζ > η . (185)

Proof. Section 5.5.

An immediate consequence of Claim 25 is that in the RSW allocation, the type-2 seller

retains all of asset 2 (resp., sells all of asset 1) if ν is positive (resp., negative):

Corollary 3. If

ν > 0, (186)

then

V 1 < ∆1 (187)

and type 2 retains all of asset 2 in the RSW allocation:

q2 =

(V 1

∆1,0

)=

(V 1−V 2

δ∆11

,0

). (188)

91

Page 92:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

If instead

ν ≤ 0, (189)

then

V 1 ≥ ∆1 (190)

and type 2 sells all of asset 1 in the RSW allocation:30

q2 =

(1,|ν |δ∆1

2

). (191)

Proof. Under (186), equation (187) follows from (183), while (188) is implied by (171),

(172), (181), (182), and (187). If instead (189) holds, then (190) follows from (183), while

(191) results from by (181), (182), and (173).

We next consider elements of S in which each type of seller gets her RSW payoff. We

show that any such allocation must coincide with the RSW allocation o·, which is unique

by Claim 25. The nontrivial part of the proof is to show that the agent’s payoff must be

zero not only ex-ante, but also ex-post.

Claim 26. Let o· be an equilibrium allocation in which each type of seller gets her RSW

allocation payoff: V t (ot) = V t for t = 1,2. Then o· is the unique RSW allocation o·.

Proof. Section 5.5.

We now turn to elements of S in which some type of seller gets more than her RSW

payoff, a property that we write formally:

P0 For some s ∈ 1,2, V s (os)> V s.

In any such allocation, the agent strictly prefers (not) to buy from the high (low) type seller,

while the low type seller gets more than her RSW payoff:

Claim 27. In any equilibrium allocation o· with property P0, U1 (o1)< 0, U2 (o2)> 0, and

x =V 1 (o1)−V 1 > 0.

30

92

Page 93:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Proof. Section 5.5.

We now give a necessary and sufficient condition for there to exist an equilibrium

allocation o· in which the type 1 seller’s payoff V(o1) exceeds her RSW payoff V 1 by

exactly some given x > 0. We also characterize the efficient such allocation: the one that

maximizes the joint surplus

J (o·) =2

∑t=1

Πt [V t (ot)+U t (ot)]= (1−δ )

2

∑t=1

Πtqt f (t) (192)

of seller and agent.31 Let

Sx =

o· ∈ S : V 1 (o1)= V 1 + x

denote the set of equilibrium allocations o· in which the type-1 seller gets exactly x more

than her RSW payoff and let

Sx = o· ∈ Sx : ∀o· ∈ Sx,J (o·)≤ J (o·)

be the efficient frontier of Sx: the set of allocations in Sx that maximize the joint surplus J.

We first show that each element of Sx satisfies four key properties in addition to P0:

Claim 28. Fix x > 0. If Sx is nonempty then so is Sx. Moreover, any allocation o· in Sx has

the following four properties:

P1 The low type sells her entire portfolio (q1 = a) in return for the payment ρ1 = a f (1)+x.

P2 The low type’s IC constraint binds: V 1 (o2)=V 1 (o1).31In MT [12], the agent is ignored in the definition of efficiency. In practice, by (158), this would restrict

us to allocations in which her expected payoff is zero. However, such allocations generally do not maximize

joint surplus. Intuitively, if the type 2 seller gets more than her RSW payoff, we can have her transfer

x = ε(a−q2

)more shares to the agent (for some small ε > 0) in return for an additional payment equal

to type 1’s opportunity cost δx f (1). This raises joint surplus without tempting the type 1 seller to imitate;

however, it also raises the agent’s payoff and thus violates MT’s notion of efficiency.

93

Page 94:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

P3 The high type either sells her whole portfolio or gets her RSW payoff: either q2 = a or

V 2 (o2)= V 2 .

P4 The high type sells all of asset 1 before selling any units of asset 2: either q21 = 1 or

q22 = 0.

We will now explicitly characterize the set Sx of efficient allocations for any given

increase x > 0 in the type 1 seller’s payoff vis-a-vis the RSW allocation. The notation

(ρ t ,qt) will refer to the outcome in a given allocation o· that corresponds to type t = 1,2.

As property P4 pins down the outcome(ρ1,q1) of the type 1 seller, outcomes in Sx can

differ only with respect to the type 2 seller’s outcome(ρ2,q2).

We now turn to the formal results that underpin Figures 1 and 2. We first consider the

subset

Spoolx =

o· ∈ Sx : q2 = q1 = a

of efficient allocations in which each type sells their entire portfolio. By P2, this implies

equal transfers as well: ρ1 = ρ2. Hence, by P1, Spoolx is the set of pooling outcomes.

Claim 29. Fix x > 0. The set Spoolx is nonempty if and only if32

Π2 ≥ ζ (193)

and x lies in the interval [γ,Π2a∆

1] (194)

(which is nonempty by (193)). In this case, Spoolx contains a single allocation, in which the

agent pays ρ t = a f (1)+ x to each type t of seller in return for his whole portfolio.

Proof. Section 5.5.

We now turn to the set

Ssepx =

o· ∈ Sx : q2 6= q1 = a

32The notation ζ is defined in (167).

94

Page 95:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

of efficient allocations for which q2 6= q1: the separating allocations. In such an allocation,

P3 implies that the high type gets her RSW payoff: V 2 (o2)= ρ2−δq2 f (2) = V 2, whence

ρ2 = δq2 f (2)+V 2. (195)

Hence it suffices to solve for q2, with the corresponding transfer ρ2 given by (195). We

do so in the following claim. Cases 1(a) and 1(b) correspond, respectively, to the lower

and upper regions of separating equilibria in Figure 1, while case 2(a) relates to the single

region of separating equilibria in Figure 2.

Proposition 5. With respect to Ssepx , there are two cases.

1. Case 1. Suppose ν > 0: in the RSW allocation, the type-2 seller retains all of asset

2. There are three subcases.

(a) Case 1(a). If x lies in the interval

(0,ν ] (196)

then Ssepx is nonempty if and only if33

Π2 ∈ [η ,1] , (197)

which has positive measure by (154): the probability of the high type cannot be

too low. In this case, Ssepx contains a single allocation, in which the type 2 seller

retains all of asset 2:

q2 =

(V 1−V 2 + x

δ∆11

,0

). (198)

(b) Case 1(b). If

x ∈ [ν ,γ] , (199)

then Ssepx is nonempty if and only if

Π2 ≥ φ (x) d

= x[

∆1−V 1 +(x−ν)∆2

δ∆12

]−1

. (200)

33The notation η is defined in (166).

95

Page 96:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

The function φ (x) is continuous and increasing,34 takes values in (0,1) when

x ≥ ν , and satisfies φ (ν) = η , φ (γ) = ζ , and limx→∞ φ (x) = δ∆12/∆2 > ζ . If

(200) holds, Ssepx contains a single allocation, in which the type 2 seller sells all

of asset 1:

q2 =

(1,

x−ν

δ∆12

). (201)

Finally, the region R defined by (199) and (200) is the union of two regions. The

first, region (i), is defined by the two conditions

Π2 ∈ [η ,ζ ] (202)

and

x ∈[ν ,φ−1 (

Π2)] , (203)

where

φ−1 (

Π2)=

[δ∆1

2

(∆1−V 1

)−ν∆2

]Π2

δ∆12−∆2Π2 . (204)

The second, region (ii), is defined by (199) and

Π2 ∈ [ζ ,1] . (205)

(c) Case 1(c). If x exceeds the upper endpoint of (199) then Ssepx is empty.

2. Case 2. Suppose ν ≤ 0: in the RSW allocation, the type 2 seller sells all of asset 1.

There are two subcases.

(a) Case 2(a). If x lies in the interval

[0,γ] , (206)

then Ssepx is nonempty if and only if

Π2 ≥ ζ , (207)

34Intuitively, φ (x) is increasing in xsince, in order for the agent to participate when the type-1 agent’s

payoff increment x over her RSW payoff is higher, the agent must believe that the agent is more likely to be

of type 2.

96

Page 97:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

in which case Ssepx contains a single allocation, in which the quantities q2 sold

by type 2 are given by (201).

(b) Case 2(b). If x exceeds the upper endpoint of (206) then Ssepx is empty.

Proof. Section 5.5.

Finally, we characterize the set

S = o· ∈ S : ∀o· ∈ S,J (o·)≤ J (o·)

of globally efficient equilibrium allocations: those equilibrium allocations that give the

highest social welfare for any payoff increment x. This set consists of the pooling equilibria

depicted in Figures 1 and 2, together with the upper, curved boundary of the region of

separating equilibria for Π2 between η and ζ in Figure 1.

Claim 30. There are three cases.

1. If Π2 ≥ ζ , S consists of those pooling allocations in which, for some x in the interval

(194), the agent pays ρ t = a f (1)+ x to each type of seller for her whole portfolio.

2. If ν > 0 and Π2 ∈ [η ,ζ ), S is a singleton consisting of the allocation in which q2 is

given by (201) for x = φ−1 (Π2).3. In all other cases, S is empty.35

Proof. Section 5.5.

5.4 The MT Approach and Security Design

We now apply the results of section 5.3 to the problem of security design. Assume an issuer

owns a stochastic cash flow Y that takes three possible values: y2 > y1 > y0 = 0. Suppose

the issuer’s beliefs about Y , conditional on his type t, satisfy the Hazard Rate Ordering

35As is Spoolx and Ssep

x for all x > 0.

97

Page 98:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

(HRO) property (DFJ, section 2.5). Assume the issuer can sell any monotone security S:

any nondecreasing function S : 0,y1,y2 → ℜ such that, for each possible Y , S (Y ) and

Y −S (Y ) are both in [0,Y ].

How do we apply MT’s approach in this setting? In section 5.2 we defined an outcome

as a pair (q,ρ) consisting of a payment ρ from agent to seller in return for the vector

q = (q1, ...,qn) of quantities of the seller’s assets. In the present setting, the vector q is

replaced by a monotone security S; let O be the set of outcomes (S,ρ). The rest is as in

section 5.2, with the set O of outcomes replaced by O. Let us call this the MTSD game (for

MT - security design).

How do we find the set of equilibria of the MTSD game? As shown in DFJ section 3.3,

a monotone security S is equivalent to a portfolio of assets, where the set of assets is given

by the maximal tranching of the cash flow Y into prioritized, monotone securities. In the

present case with three possible outcomes of the cash flow, this maximal tranching consists

of the two securities F1 (Y ) = y11Y∈y1,y2 and F2 (Y ) = (y2− y1)1Y=y2 . Let MTAS refer to

the 2x2 MT asset sale game in which these are the two assets that the seller wishes to sell.

As in DFJ section 3.3, there is an isomorphism between the MTSD and MTAS, which

is as follows. For any security S in the MTSD game, the quantity of asset i = 1,2 in the

MTAS game is given by qSi = [S (yi)−S (yi−1)]/(yi− yi−1). As shown in DFJ section 3.3,

the monotone security S is payoff-equivalent to the portfolio qS =(qS

1,qS2)

of assets F1 and

F2: for any realization Y of the cash flow, the security S and the portfolio qS imply the same

ex-post transfer from issuer to agent. And any portfolio q = (q1,q2) of the assets F1 and F2

in the MTAS game is payoff-equivalent to the monotone security Sq = qF (Y ) in the MTSD

game, where F (Y ) is the vector (F1 (Y ) ,F2 (Y )) of realized payouts of assets F1 and F2.

We now solve the MTSD game by applying the results of section 5.3 to the MTAS

game. For each type t = 1,2 and each asset i = 1,2, let fi (t) = E [Fi (Y ) |t] denote the

expected payout of asset Fi when the issuer’ type is t. Let f (t) = ( f1 (t) , f2 (t)) denote

the vector of these expected payouts for type t. The function f satisfies (154) since

HRO implies FOSD and since the assets are monotone functions of Y . It satisfies (155)

98

Page 99:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

by DFJ’s Proposition 5 and by HRO. Consequently, the set of equilibrium allocations((q1,ρ1) ,(q2,ρ2)) identified in section 5.3 in the MTAS game for this function f is

isomorphic to the set of equilibrium allocations((

S1,ρ1) ,(S2,ρ

2)) of the MTSD game:

for each type t = 1,2, the agent pays the issuer ρ t in return for the monotone security

St (Y ) = qtF (Y ).

Using this equivalence, we can now discuss the main implications of section 5.3 for

the MTSD game. As in section 5.3, we restrict to allocations that are efficient contingent

on giving the low type a variable amount x more than her RSW payoff. By Claim 25, the

RSW allocation is always an equilibrium of the MTAS game. What its equivalent in the

MTSD game? By (168), in the RSW allocation, the low type sells her whole portfolio.

This is equivalent, in the MTSD game, to standard debt (secured by Y ) with face value y2.

By (171) and (172), the type 2 issuer uses a hurdle class strategy (DFJ section 3.4) in the

MTAS game: either q22 is zero (whence the hurdle class c is 1) or q2

1 is one (whence c = 2).

In the first case, the corresponding security in the MTSD game is

q2F (Y ) = q21y11Y∈y1,y2

which is just standard debt with face value q21y1. In the second case, the security sold by

the high type in the MTSD game is

q2F (Y ) = y11Y∈y1,y2+ q22 (y2− y1)1Y=y2

which is just standard debt with face value y1 + q22 (y2− y1). Indeed, this is just the result

of DFJ section 3.4 as the RSW and RLP allocations coincide (Claim 18).

We now turn to efficient equilibrium allocations other than RSW.36 By Claim 28, in

each such allocation the low type sells her whole portfolio which, as shown above, is

equivalent to standard debt with face value y2. As for the high type, Claim 28 implies

that he uses a hurdle class strategy which, as shown in DFJ section 3.4, is also equivalent

to standard debt. Since the low type chooses the highest possible face value, the qualitative

36By Proposition 4, these do not satisfy the intuitive criterion.

99

Page 100:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

results of DFJ section 3.4 carry over to the MTSD game: the issuer sells standard debt with

a face value that is nonincreasing in his type.37

How do these non-RSW allocations differ from the RSW allocation? By Claim 27, the

low type is paid some x> 0 more than the fair value of her security. (In the RSW allocation,

she is paid exactly this fair value.) This loosens her IC constraint which, in turn, enables

the high type to raise the face value of the security he sells relative to the outcome of DFJ

section 3.4. In the case in which Π2 exceeds the constant ζ defined in (167), the high type

can raise his face value all the way to the face value y2 chosen by the low type: the two

types can pool, selling the same standard debt security (By Claim 29). As in section 5.3,

neither the pooling outcomes nor the non-RSW separating outcomes survive the intuitive

criterion.

5.5 MT Approach: Proofs

PROOF OF CLAIM 15. Let mΣ denote the DRM o·Σ

defined in (133).38 Further, consider

the profile Σ in which the principal always chooses the mechanism mΣ, the agent accepts,

and the principal chooses the outcome otΣ

corresponding to his type; if any player ever

deviates, they revert to the original equilibrium Σ.39 If Σ is an equilibrium, then it is a

DRM equilibrium that implements o·Σ. We next show that no player ever wants to deviate

in Σ, working backwards.

Stage 2. Suppose there has been no prior deviation: the principal offered mΣ in stage 1.

37More precisely, it is decreasing in his type in the separating equilibria and constant in his type in the

pooling outcomes.

38That is, if the agent accepts and the principal reports ”t”, the mechanism mΣ implements the outcome otΣ.

39Formally, Σ =(

p··,Π· (·) , π·

)where, for all t, pt

mΣ= 1 and, for m 6= mΣ, pt

m = 0. If m = mΣ, the agent’s

beliefs are just the prior beliefs(Π· (mΣ) = Π·) and the stage-2 distribution πmΣputs unit weight on the action

profile(sP· ,s

A)= ((0, ...,T ) ,”accept”). (By sP

· = (0, ...,T ) we mean the pure strategy in stage 2 in which

the principal of each type t truthfully reports his type.) If m 6= mΣ, beliefs and stage-2 play are as in Σ:

Π· (m) = Π· (m) and πm = πm.

100

Page 101:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

If the agent chooses ”reject”, she gets zero. Let MΣ ⊂M denote the set of mechanisms that

some type t chooses in Σ with probability ptm > 0. If the agent accepts, she gets

T

∑t=0

ΠtU t (ot

Σ

)=

T

∑t=0

Πt

∑m∈M

ptm ∑

sP· ∈(SP

m)T+1

∑sA∈SA

m

πm

(sP· ,s

A)

U t(

osPt ,s

A

m

)by (121) and (133)

= ∑m∈MΣ

T

∑s=0

Πs ps

m

T

∑t=0

Πt ptm

∑Ts=0 Πs ps

m∑

sP· ∈(SP

m)T+1

∑sA∈SA

m

πm

(sP· ,s

A)

U t(

osPt ,s

A

m

)

= ∑m∈MΣ

T

∑s=0

Πs ps

m ∑sP· ∈(SP

m)T+1

∑sA∈SA

m

πm

(sP· ,s

A) T

∑t=0

Πt (m)U t

(osP

t ,sA

m

)

= ∑m∈MΣ

T

∑s=0

Πs ps

m ∑sP· ∈(SP

m)T+1

∑sA∈SA

m

πm

(sP· ,s

A)

uAm

(sP· ,s

A)

by (127)

= ∑m∈MΣ

T

∑s=0

Πs ps

mUm (πm) by (130)

which is nonnegative by Claim 14: she is willing to accept mΣ. As for a type-t principal,

in Σ he has the option of imitating any other type t ′ yet chooses not to. Thus, his payoff

from doing so, V t(

ot ′Σ

), must not exceed the payoff he gets in equilibrium in Σ, which is

V t (otΣ

). Since these are also the principal’s payoffs from choosing t ′ and t, respectively, in

Σ, he will not deviate in Σ either.

Stage 1. If the principal deviates in stage 1 by proposing some m 6= mΣ, then he and

the agent will conform to Σ thereafter. Thus, his payoffs in Σ and Σ from proposing m are

the same. But this payoff in Σ cannot exceed his payoff from sticking to his equilibrium

strategy in Σ, which is V t (otΣ

): the same payoff that he gets in Σ in equilibrium. Accordingly,

the principal is willing not to deviate in stage 1.

Finally, beliefs in an equilibrium are arbitrary after deviations, so the agent’s beliefs

following an offer m 6= mΣ may equal Π· (m) or anything else. We conclude that Σ is an

equilibrium as claimed. Q.E.D.Claim 15

PROOF OF CLAIM 16. The payoff functions V t and U t are linear in the components qt and

ρ t of o·. Hence any convex combination of any two allocations (o·, o·) that each satisfies

(136) and (137) must also satisfy these conditions. Thus, the set of allocations that satisfy

101

Page 102:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

(136) and (137) is convex. The result then follows from Proposition 16.E.2 in Mas-Collel,

Whinston, and Greene [11]. Q.E.D.Claim 16

PROOF OF CLAIM 18.

Part 1. Define the following set.

Definition 18. AUp is the set of allocations o· ∈O that satisfy upwards incentive-compatibility

for the principal:

V s (os)≥V s (ot) for all types 0≤ s < t ≤ T ; (208)

and, for all types, are individually rational for the agent:

U t (ot)≥ 0 for all types 0≤ t ≤ T . (209)

The proof of Part 1 consists of showing (a) any RSW allocation is Pareto optimal in

AUp and (b) any solution to RLP is Pareto optimal in AUp and vice-versa. We begin with

two useful properties of AUp.

Claim 31. AUp contains a Pareto optimal allocation.

Proof. For any allocation o·, let SW (o·) denote the sum ∑Tt=0V t (ot) of payoffs of the

different types of principal. The function SW (·) is linear in the arguments qt and ρ t of

o· and thus continuous in o·. Define AUp0 to be the subset of AUp for which each type of

principal gets a nonnegative payoff:

V t (ot)≥ 0 for all types t. (210)

The set AUp0 is nonempty (it includes o·0), closed (it is defined by weak inequalities of linear

functions), and bounded (as the sum of the payoffs U t (ot)+V t (ot) = (1−δ )qotf (t) is

bounded and, by (209) and (210), each payoff is nonnegative). Hence AUp0 is compact. As

SW (·) is linear and thus continuous, it must have a maximand o· on the nonempty, compact

set AUp0 by the extreme value theorem. Clearly, o· is Pareto optimal in AUp

0 . If there is an

allocation o· in AUp−AUp0 that Pareto dominates o· then V t (ot)≥V t (ot)≥ 0 for each type

t whenceo· is in AUp0 - a contradiction.40

40See Mas-Collel, Whinston, and Greene [11, Prop. 16.E.2].

102

Page 103:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Claim 32. If o· ∈ AUp is not Pareto optimal in AUp, then there exists an allocation o· in AUp

that Pareto dominates o· and is Pareto optimal in AUp.

Proof. By definition 5, there is an allocation o· ∈ AUp that Pareto dominates o·. Let

AUpo· =

o· ∈ AUp : V t (o·)≥V t (o·) for each type t

(211)

be the set of allocations in AUp that are at least as good as o· is for each type t of principal.

The set AUpo· is nonempty (as it contains o·); closed (as it is defined by weak inequalities of

linear functions); and bounded (as the sum of the payoffs U t (ot)+V t (ot) = (1−δ )qotf (t)

is bounded, U t (ot)≥ 0, and V t (ot)≥V t (o·)>−∞). Thus, the function SW (o·)=∑Tt=0V t (ot)

on AUpo· has a maximizer o· in AUp

o· , which is Pareto optimal in AUpo· and thus also in AUp.

(An allocation o· that Pareto dominated o· in AUp would, by transitivity, Pareto dominate

o· and thus lie in AUpo· , contradicting the Pareto optimality of o· in AUp

o· .) Finally, as o· is in

AUpo· , it Pareto dominates o· since o· does.

We next show that the solutions to RLP coincide with the Pareto optimal allocations in

AUp. More precisely, for each Pareto optimal allocation o· in AUp and each type t, the payoff

V t (ot) of the principal equals the seller’s unique payoff u∗ (t) in RLP, while the agent’s

payoff U t (ot) is zero as in RLP. Moreover, the issuance function defined by q∗ (t) = qot

solves RLP. Finally, the expected transfer ρotcoincides with the unique transfer function

ρ∗ (t) in RLP and equals the conditional expected value q∗ (t) f (t) of the issuance in RLP.

The Claim also proves the converse: any solution to RLP corresponds to a Pareto-optimal

element of AUp.

Claim 33. 1. Suppose o· is Pareto optimal in the set AUp. Then for each type t, the

agent’s payoff U t (ot) is identically zero and the functions given by u∗ (t) = V t (ot),

ρ∗ (t) = ρot, and q∗ (t) = qot

solve RLP.

2. Let (u∗,ρ∗,q∗) solve RLP. Then any allocation o· that satisfies qot= q∗ (t) and ρot

=

ρ∗ (t) for all types t is Pareto optimal in the set AUp and, for each type t, yields the

payoffs V t (ot) = u∗ (t) and U t (ot) = 0.

103

Page 104:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Proof. Part 1. Suppose o· is Pareto optimal in AUp. Let t be the lowest type for which

U t (ot) is not identically zero:

0 <U t (ot)= qotf (t)−ρ

ot. (212)

If the IC constraint (208) does not bind for any s < t, then we can raise the transfer ρotthat

type t gets by ι =U t (ot)> 0 without violating (208) or (209), which contradicts the Pareto

optimality ofo· in AUp. Now suppose the IC constraint (208) does bind for some type s < t.

Let

w = min

s≤ t : V s (os) =V s (ot) and U s (ot)> 0

(213)

be the lowest type s ≤ t for which (208) binds and U s (ot) > 0. (If there is no such type

s < t, then w equals t.) By construction, Uw (ot) is positive. Let

ε ∈(

0,min

1,Uw (ot)

2(1−δ )qot f (w)

)(214)

and define41

ι (ε) = min

Uw (ot)

2,min

s<w

(V s (os)−V s (ot)+ εδqot

[ f (w)− f (s)])

> 0. (215)

Now consider the alternative allocation o· given by os = os for s 6= w and ow = (q,ρ) where

q = (1− ε)qotand ρ = ρot − εδqot

f (w)+ ι (ε). For s < w,

V s (os)−V s (ow) =V s (os)−V s (q,ρ)

=V s (os)−[ρ

ot− εδqot

f (w)+ ι (ε)−δ (1− ε)qotf (s)

]=V s (os)−V s (ot)+ εδqot

[ f (w)− f (s)]− ι (ε)

which is nonnegative by (215): no type s < w prefers ow to os. Thus, the IC constraint

(208) is satisfied in o·. Moreover,

Uw (ow) = (1− ε)qotf (w)−

ot− εδqot

f (w)+ ι (ε)]

=Uw (ot)− ε (1−δ )qotf (w)− ι (ε)

41If w = 0, then ι is simply Uw (ot)/2 > 0. If w > 0, then the inner min in (215) is positive since for each

s < w, either V s (os)>V s (ot) or U s (ot)≤ 0; the latter and Uw (ot)> 0 jointly imply 0 <Uw (ot)−U s (ot) =

qot[ f (w)− f (s)].

104

Page 105:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

which is nonnegative by (214) and (215): the IR constraint (209) for type w is also satisfied

in o·. Hence, o· lies in AUp. But the payoff of the type-w principal in o· is

V w (ow) =[ρ

ot− εδqot

f (w)+ ι (ε)]−δ (1− ε)qot

f (w)

=V w (ot)+ ι (ε)>V w (ot)which contradicts the Pareto optimality of o· in AUp. We conclude that, if o· is Pareto

optimal, then U t (ot) = 0 for all t.

Now let t be the lowest type for which the functions u∗ (t) = V t (ot), ρ∗ (t) = ρot,

and q∗ (t) = qotdo not solve RLP. As U t (ot) = 0, it follows that ρot

= qotf (t), whence

V t (ot) = (1−δ )qotf (t). Hence, if q∗ (t) = qot

solves RLP, so do u∗ (t) = V t (ot) and

ρ∗ (t) = qotf (t). Thus, q∗ (t) = qot

must not solve RLP. By assumption, t is the lowest

type for which the conclusion fails: u∗ (s) = V s (os), ρ∗ (s) = ρos, and q∗ (s) = qos

solve

RLP for each s < t. As o· is Pareto optimal in the set AUp, we have, for all 0 ≤ s < t,

u∗ (s) = V s (os) ≥ V s (ot) = qot[ f (t)−δ f (s)] whence qot

is in Qt . Thus, there must be

a q′ in Qt such that (1−δ )q′ f (t) > (1−δ )qotf (t). Let us now replace the allocation o·

with the allocation o· given by ot = (q′,q′ f (t))and, for all s 6= t, os = os. As U t (ot) = 0,

o· satisfies (209) for t = T . Since q′ is in Qt , for each s < t we have V s (os) = u∗ (s) ≥

q′ [ f (t)−δ f (s)] =V s (ot): o· satisfies (208) for t = T . Thus, o· is in AUp, each type s 6= t

of principal is indifferent between o· and o·, and the type t principal is better off under o·

than o·. This contradicts the Pareto optimality of o· in AUp.

Part 2. Let (u∗,ρ∗,q∗) solve RLP. Let o· be an allocation that satisfies qot= q∗ (t) and

ρot= ρ∗ (t) for all types t. This implies that

V t (ot)= ρot−δqot

f (t) = ρ∗ (t)−δq∗ (t) f (t) = u∗ (t) (216)

(which is positive by PROPOSITION 1) and

U t (ot)= qotf (t)−ρ

ot= q∗ (t) f (t)−ρ

∗ (t) = 0

105

Page 106:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

as claimed. Moreover, for s < t,

V s (os) = u∗ (s) by (216)

≥ ρ∗ (t)−δq∗ (t) f (s) as q∗ (t) ∈ Qt and ρ

∗ (t) = q∗ (t) f (t)

= ρot−δqot

f (s) =V s (ot) by hypothesis on o·.

It follows that o· lies in AUp. By Claim 31, AUp contains a Pareto optimal element o·.

By part 1, for each type t, the agent’s payoff U t (ot) is zero and the functions given by

u∗ (t) = V t (ot), ρ∗ (t) = ρ ot, and q∗ (t) = qot

solve RLP. But by PROPOSITION 1, the

solutions u∗ (·) and ρ∗ (·) to RLP are unique. Accordingly, V t (ot) = V t (ot) for each t,

whence o· is Pareto optimal in AUp.

AUp has the following relation to WIE:

Claim 34. 1. If o· is Pareto optimal in AUp, then it is WIE.

(a) If o· is WIE and U t (ot)≥ 0 for each type t, then o· is in AUp.

(b) If o· is WIE and U t (ot) = 0 for each type t, then o· is Pareto optimal in AUp.

Proof. Part 1. By part 1 of Claim 33, the functions u∗ (t) = V t (ot), ρ∗ (t) = ρot, and

q∗ (t) = qotsolve RLP. Hence, by PROPOSITION 1, for any types s and t,42

V t (ot)= u∗ (t)≥ ρ∗ (s)−δq∗ (s) f (t) = ρ

os−δqos

f (t) =V s (ot) . (217)

Thus, o· satisfies (134). Now suppose that o· is not WIE. Then there is an allocation o·

that satisfies (135), (136), and (137). By (136), o· satisfies (208). Since o· is in AUp,

V t (ot) ≥ V t (ot) by (135) and U t (ot) ≥U t (ot) = 0 by (137): o· satisfies (209). Thus, o·

is in AUp and, by (135), Pareto dominates o·, which contradicts the assumption that o· is

Pareto optimal in AUp.

42PROPOSITION 1 characterizes a profile e∗, in which a type t seller sells the quantities q∗ (t) in return

for the revenue ρ∗ (t), and gets the payoff u∗ (t). Part 3 shows that e∗ is an equilibrium, in part by showing

the inequality in (217) holds for all types s and t.

106

Page 107:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Part 2(a). Since U t (ot) ≥ 0 for each type t, o· satisfies (209). Since o· is WIE, it also

satisfies (208) by (134): it is in AUp.

Part 2(b). By part 2(a), o· is in AUp. If o· is not Pareto optimal in AUp, then by Claim

32 there exists an allocation o· in AUp that Pareto dominates o· and is Pareto optimal in

AUp. Hence, by part 1 of Claim 33, the functions given by u∗ (t) =V t (ot), ρ∗ (t) = ρ ot, and

q∗ (t) = qotsolve RLP. Thus, by part 1 of PROPOSITION 1, the allocation o· is incentive

compatible: it satisfies (136). As o· is in AUp, it also satisfies (137) by (209). Finally,

by assumption, o· Pareto dominates o·: o· = o· satisfies (135). But this contradicts the

assumption that o· is WIE.

Part 1 of Claim 18 now follows from Claims 33 and the following result.

Claim 35. Any RSW allocation o· is Pareto optimal in AUp.

Proof. Let o· be an RSW allocation. By Claim 31, there is an allocation o· that is Pareto

optimal in AUp and, by part 1 of Claim 33, the functions given by u∗ (t) =V t (ot), ρ∗ (t) =

ρot, and q∗ (t) = qot

solve RLP. Thus, by part 1 of PROPOSITION 1, the agent’s payoff

U t (ot) is identically zero: o· satisfies (140). By part 3 of PROPOSITION 1, o· also satisfies

(139). Thus, by definition 14, V t (ot) is not less than V t (ot) which, we have shown, equals

u∗ (t). Since, moreover, o· satisfies (208) by (134) and (209) by (140), o· is in AUp.Now

suppose o· is not Pareto optimal in AUp. By Claim 32, there exists an allocation o· in AUp

that Pareto dominates o· and is Pareto optimal in AUp. Hence, o· satisfies (139) by part 1 of

Claim 34 and (140) by part 1 of Claim 33, so V t (ot)≥V t (ot) for all t by Definition 14: o·

does not Pareto dominate o· - a contradiction.

Part 2.An RSW allocation o· exists by Claim 17 and, by part 1, satisfies

V t =V t (ot)= u∗ (t) . (218)

Now consider the alternative allocation o· = (q·,ρ ·) defined by qt = q∗ (t) and ρ t = ρ∗ (t).

We claim that it is an RSW allocation as well. We first verify that it satisfies (139) and

107

Page 108:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

(140) for all t. Equation (140) holds since, for each s, U s (os) equals q∗ (s) f (s)−ρ∗ (s)

which, by (143), is zero: the agent’s reservation payoff. Moreover,

V s (os) = ρ∗ (s)−δq∗ (s) f (s) = u∗ (s) (219)

for all s by (144) and

V s (or) = ρ∗ (r)−δq∗ (r) f (s) = q∗ (r) [ f (r)−δ f (s)]

by (143). Hence, to show (139), it suffices to show that u∗ (s) ≥ q∗ (r) [ f (r)−δ f (s)] for

all types r,s. But this has already been shown in the proof of part 3 of PROPOSITION 1.

We conclude that o· solves (139) and (140) for all t. Finally, for each t, V t (ot) attains the

maximum payoff V t possible in Program IIt by (218) and (219). Hence, ot solves Program

IIt for all t: o· is an RSW allocation. Q.E.D.Claim 18

PROOF OF CLAIM 19. The payoff functions V t and U t are linear in qt and ρ t . Hence

any convex combination of any two allocations (o·, o·) that each satisfies (147) and (148)

must also satisfy these conditions. Thus, the set of allocations that satisfy (147) and (148)

is convex. The result then follows from Proposition 16.E.2 in Mas-Collel, Whinston, and

Greene [11]. Q.E.D.Claim 19

PROOF OF CLAIM 20. Let C = o· ∈ A : V t (ot)≥V t (os) for all s, t denote the set of

incentive-compatible allocations. Since the function V t (o) is linear in each element of

o = (qo,ρo), C is a convex subset of A.43 Let o· be WIE. By Claim 16 and Bertsekas [3,

Proposition 5.3.1],44

o· ∈ argmaxo·∈C

Tt=0 wtV t (ot)+∑

Tt=0 ν

t [U t (ot)−U t (ot)]43Defined in (123), A is the set of all possible allocations.

44To satisfy condition (1) in Bertsekas [3, Proposition 5.3.1], begin with o· and then lower the transfer ρ t

that the agent pays the principal by some common (type-independent) amount ε > 0. The resulting allocation

is in C and, moreover, strictly satisfies (137): condition (1) holds.

108

Page 109:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

for some positive weights wtTt=0 and nonnegative multipliers ν tT

t=0. Let ν = ∑Tt=0 ν t .

For each t, let Πt equal vt/ν if ν > 0 and 1/(T +1) otherwise. Then

o· ∈ argmaxo·∈C

Tt=0 wtV t (ot)+ν ∑

Tt=0 Π

t [U t (ot)−U t (ot)] .Hence, by Claim 19 and Bertsekas [3, Proposition 5.3.1], o· is IE with respect to the beliefs

Π·: Π(o·) is nonempty. The proof of convexity of the set Π(o·) is identical to that of MT

[12, p. 17]. Q.E.D.Claim 20

PROOF OF CLAIM 21. Let o· be an RSW allocation. By Claim 17, o· is Pareto optimal

in AUp whence, by part 1 of Claim 33, the agent’s payoff U t (ot) is identically zero and

the functions given by u∗ (t) = V t (ot), ρ∗ (t) = ρ ot, and q∗ (t) = qot

solve RLP. Now fix

a small ε > 0 and consider the allocation o· given by, for each type t, qt = qotand ρ t =

ρ ot − (t +1)ε . Then for any types s and t,

V s (ot)=V s (ot)− (t +1)ε. (220)

If type s proposes o·, the agent will accept. Why? First, the agent gets a positive payoff if

type s chooses os as

U s (os) = (s+1)ε > 0. (221)

Moreover, type s will not imitate any type t > s since V s (os)−V s (ot) =V s (os)−V s (ot)+

(t− s)ε by (220) which is positive as o· is in AUp. Finally, if type t imitates type s < t then,

by ASSUMPTION A, the agent gets no less than U s (os) which again is positive by (221).

As the agent knows her payoff will be positive,she will accept. As for the principal, if her

type is s then by proposing o· and then choosing os she receives V s (os)−(s+1)ε by (220).

Thus, she will deviate from any putative equilibrium that offers her less than this. As this is

true for all ε > 0,her payoff in any equilibrium cannot be less than V s (os). Q.E.D.Claim 21

PROOF OF PROPOSITION 3.

By Claim 23, the RSW allocation o· is IE relative to some nonnegative beliefs Π· ∈

ℜT+1+

109

Page 110:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Part 1: only if. Condition (151) is necessary since, otherwise, the principal could

strictly gain by imitating a different type. Condition (152) is necessary since, otherwise,

the agent could do better by always playing ”reject”. And (153) is necessary by Claim 21.

We conclude that (151), (152), and (153) are necessary conditions for o· to be the expected

allocation of an equilibrium.

Part 2: if. Consider any allocation o·= o· that satisfies (151), (152), and (153). We will

prove the existence of a DRM equilibrium o·, which is intuitive if o· is an RSW allocation

o·.

In such an equilibrium, the principal always chooses the DRM o·. Thus, if the principal

conforms to the equilibrium in stage 1, the agent’s interim beliefs are simply her prior

beliefs Π·: by (152), she will accept in stage 2. Moreover, the principal is willing to

truthfully report his type in the DRM o· by (151). It remains only to show that there exist

interim beliefs that deter the principal from deviating to a mechanism other than o·, and

that these beliefs are intuitive if o· is an RSW allocation.

For any alternative mechanism m ∈M, define

∆Σm =

Π· ∈ ∆T : Πt = 0 for all t ∈ Jm

if Jm 6= ϒ

∆T if Jm = ϒ

(222)

where Jm, defined in (149), is the set of types t who will never choose m over the RSW

allocation. If some types are not in Jm, then ∆Σm is the set of interim beliefs that assign

probability zero to all types in Jm; else it is the full set of all possible beliefs. As m is

unexpected, any interim beliefs Π· (m) are consistent with Bayes’s Rule. If, in addition, the

interim beliefs Π· (m) lie in ∆Σm, then they satisfy IntΣm:

Claim 36. ∆Σm is compact and the beliefs in ∆Σ

m satisfy IntΣm.

Proof. Compactness is immediate from (222). As for IntΣm, if Jm 6= ϒ then the beliefs in ∆Σm

put zero weight on any type in Jm by (222). And if Jm = ϒ, then no type t in ϒ−Jm satisfies

(150) as ϒ−Jm is empty: IntΣm holds for any beliefs. In both cases, IntΣm holds for all beliefs

in ∆Σm.

110

Page 111:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

We will show the following condition holds for each m.

Deter(m). For each mechanism m, there exist interim beliefs Π· (m) in ∆Σm and associated

equilibrium payoffs(V · (m) ,U (m)

)in Ψm

(Π· (m)

)such that V t (m) ≤ V t for all t:

the type-t principal does not prefer m to ot .

Suppose Deter(m) holds for each out-of-equilibrium mechanism m. Consider the strategy

profile in which any unexpected mechanism m leads to the beliefs Π· (m) in ∆Σm and associated

equilibrium payoffs(V · (m) ,U (m)

)described in Deter(m). Then as long as the agent will

surely accept the DRM o·, a type-t principal is willing to choose o· and then, by (151), to

choose the outcome ot . And since each type of principal chooses o·, the agent’s interim

beliefs in response to o· are her prior beliefs Π·, so she is indeed willing to accept by (152).

Hence, the DRM o· is supported by an equilibrium. If, further, o· = o·, then IntΣm holds

since Π· (m) lies in ∆Σm: the given equilibrium is intuitive, so the RSW allocation o· is the

expected allocation of an intuitive equilibrium as claimed.

It remains only to prove Deter(m) for each mechanism m other than the DRM o·.

Suppose, to the contrary, that there is a mechanism m for which Deter(m) fails to hold: for

any interim beliefs Π· (m) in ∆Σm and equilibrium payoffs

(V · (m) ,U (m)

)in Ψm

(Π· (m)

),

there is a type t for whom V t (m)> V t . We will derive a contradiction. A key step is the use

of Kakutani’s [9] fixed point theorem. To verify the assumptions of Kakutani’s theorem,

we must first derive a series of technical results.

Claim 37. Let S be any compact subset of ∆T . The correspondences φm : S⇒ Φm and

ψm : S⇒ ℜT+1×ℜ defined in (129) and (132) are upper hemicontinuous and, for any

interim beliefs Π· ∈ S, the sets φm

(Π·)

and ψm

(Π·)

are nonempty and convex.

Proof. We first show an equivalence between CEIGΠ·m and Aumann’s [2] original notion of

correlated equilibrium in a symmetric information setting.

Lemma 7. A distribution πm∈Φm is a CEIGΠ·m if and only if it is a correlated equilibrium

(Aumann [2]) of the complete information game with pure strategy sets(SP

m)T+1 and SA

m

111

Page 112:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

and payoff functions

uPm

(sP· ,s

A)=

T

∑t=0

V t(

osPt ,s

A

m

). (223)

and uAm,Π·

for the principal and agent, respectively. These payoff functions are continuous

in the interim beliefs Π·, and the pure strategy sets(SP)T+1 and SA are finite.

Proof. Equation (125) holds if and only if

∑sA∈SA

m

πm

(sP· ,s

A)

uPm

(sP· ,s

A)≥ ∑

sA∈SAm

πm

(sP· ,s

A)

uPm

(sP· ,s

A)

for any sP· , s

P· ∈

(SP

m)T+1

(224)

Why? If (125) holds then, summing over types t, one obtains (224). Conversely, if (125)

does not hold then, for some type t and actions sP· ,s

P· , ∑sA∈SA

mπm(sP· ,s

A)V t(

osPt ,s

A

m

)<

∑sA∈SAm

πm(sP· ,s

A)V t(

osPt ,s

A

m

). Defining sP

· by sPt = sP

t and, for each t ′ 6= t, sPt ′ = sP

t ′ , we have

∑sA∈SA

m

πm

(sP· ,s

A)

uPm

(sP· ,s

A)< ∑

sA∈SAm

πm

(sP· ,s

A)

uPm

(sP· ,s

A),

whence (224) fails, as claimed. Hence, πm is a CEIGΠ·m if and only the conditions (126)

and (224) hold. By Fudenberg and Tirole [7, p. 57], these two conditions are necessary

and sufficient for πm to be a correlated equilibrium in the sense of Aumann [2]. Finally,

continuity and finiteness are trivial.

By Lemma 7 and Cotter’s [5] result for the degenerate symmetric information case, the

correspondence φm is upper hemicontinuous on ∆T and, for any interim beliefs Π· ∈ ∆T ,

the set φm

(Π·)

is nonempty and convex. By (132), ψm

(Π·)

is nonempty since φm

(Π·)

is. By (130) and (131), for any λ in [0,1],

λ(V ·m (πm) ,Um (πm)

)+(1−λ )

(V ·m(π′m),Um

(π′m))

=(V ·m(λπm +(1−λ )π

′m),Um

(λπm +(1−λ )π

′m))

whence ψm

(Π·)

is convex as well.

It remains to show that φm and ψm are upper hemicontinuous on any compact set

S ⊂ ∆T . Since φm is upper hemicontinuous on ∆T , if(

Π·n,πnm

)→(

Π·,πm

)with Π·n ∈ ∆T

112

Page 113:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

and πnm ∈ φm

(Π·n

)for all n, then πm ∈ φm

(Π·)

, so it suffices to show that if Π·n ∈ S for

all n, then Π· = limn→∞ Π·n is also in S, but this must hold as S is compact: φm is upper

hemicontinuous on S. As for upper hemicontinuity of ψm on S, suppose(

Π·n,(V ·n,Un

))→(

Π·,(V ·,U

))with Π·n ∈ S and

(V ·n,Un

)∈ ψm

(Π·n

)for all n. As just shown, Π· ∈ S, so we

must merely show that(V ,U

)∈ ψm

(Π·)

. By (132), for each nthere is an πnm ∈ φm

(Π·)

such that V ·n = V ·m (πnm) and Un = Um (πm). By taking subsequences if needed, we may

assume that πm = limn→∞ πnmexists. Hence,

(Π·n,π

nm

)→(

Π·,πm

)with πn

m ∈ φm

(Π·n

)for

all n. Thus, by the upper hemicontinuity of φm on S, πm ∈ φm

(Π·)

whence, by (132),(V ·m (πm) ,Um (πm)

)∈ ψm

(Π·)

. Finally, since πnm converges to πm, (130) and (131) imply

that(V ·m (πn

m) ,Um (πnm))=(V ·n,Un

)converges to

(V ·m (πm) ,Um (πm)

)but by assumption

limn→∞

(V ·n,Un

)=(V ·,U

)whence

(V ·,U

)=(V ·m (πm) ,Um (πm)

)∈ ψm

(Π·)

as claimed:

ψm is upper hemicontinuous on S. This concludes the proof of Claim 37.

Let

Ψm =⋃

Π·∈∆T

ψm

(Π·)

(225)

be the set of all correlated equilibrium payoff vectors in the mechanism m, for any interim

beliefs Π· in ∆T .

Claim 38. Ψm is compact.

Proof. We first show that it is bounded. Since we assume ρ ∈ [−ρ,ρ] for some arbitrarily

large but finite ρ , the worst that could happen to the principal is to pay the agent ρ and also

give the agent his entire portfolio; his payoff would then be at least−ρ−δa f (T ). The best

that could occur is for him to retain his whole portfolio and receive a transfer of ρ from the

agent, giving him a payoff of ρ . Hence, the principal’s payoff lies in [−ρ−δa f (T ) ,ρ] for

each type. Likewise, the agent’s payoff lies in [−ρ,ρ +a f (T )]. Hence, Ψm is bounded.

We now show that it is closed. Let(V ·n,Un

)→(V ·,U

)with

(V ·n,Un

)∈ Ψmfor all n.

Then by definition of Ψm, for each n there is a Π·n in ∆T such that(V ·n,Un

)∈ ψm

(Π·n

).

By taking subsequences if needed, we can assume that Π·n converges to a limit Π·. That is,(Π·n,(V ·n,Un

))→(

Π·,(V ·,U

))with Π·n ∈ ∆T and

(V ·n,Un

)∈ψm

(Π·n

)for all n. As shown

113

Page 114:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

in the proof of Claim 37, this implies that Π· ∈ ∆T and(V ,U

)lies in ψm

(Π·)

which is a

subset of Ψm: Ψm is closed.

By Claim 23, the RSW allocation is IE with respect to some nonnegative prior beliefs

Π·. Let us now modify the game in two ways: (a) the agent’s prior beliefs are Π· rather

than Π· and (b) the principal must choose either the mechanism m or the RSW DRM o·. In

the modified game G, let (V ·,U) ∈Ψm denote the correlated equilibrium payoff vector that

results when the principal chooses m. In this situation,

Pt (V ·,U) = argmaxp∈[0,1]

[pV t +(1− p)V t

](226)

is the type-t principal’s set of optimal probabilities of choosing m. Let P· (V ·,U)=∏Tt=0 Pt (V ·,U)

be the corresponding set of optimal probability vectors45 and let

P(Ψm) = P· (V ·,U) : (V ·,U) ∈Ψm

be the image of Ψm under P·. Clearly, P(Ψm) is a subset of

PΣm =

p· ∈ [0,1]T+1 : pt = 0 for all t ∈ Jm

. (227)

Claim 39. The correspondence P· : Ψm⇒ PΣm is upper hemicontinuous and, for any payoff

vector (V ·,U) ∈Ψm, the set P· (V ·,U) is nonempty and convex. Moreover, PΣm is compact.

Proof. Compactness of PΣm is immediate from (227). For any V · in ℜT+1, Pt (V ·) is

clearly nonempty and convex whence so is P· (V ·). For upper hemicontinuity, suppose

that (V ·n, p·n)→ (V ·, p·)with V ·n ∈ S and p·n ∈ P· (V ·n)for all n. Then V · ∈ S as S is compact.

It remains to show that p· ∈ P· (V ·) - or, equivalently, that pt ∈ Pt (V ·) for each type t.

This is trivial if V t = V t since then Pt (V ·) = [0,1] which must contain pt . If instead

V t < (>)V t , then Pt (V ·) equals 0 (resp., 1) but also, for high enough n, V tn < (>)V t

whence ptn equals 0 (resp., 1) and thus limn→∞ pt

n is also 0 (resp., 1) which lies in Pt (V ·), a

contradiction. Since this is so for each type t, p· = limn p·n lies in P· (V ·) as claimed.

45The symbol ∏ here denotes the Cartesian product.

114

Page 115:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

In the modified game G, let m lead to the correlated equilibrium payoff vector (V ·,U).

Let the principal’s strategy be p· where pt is the type t principal’s chance of choosing m.

If the principal chooses m, what is the set of possible interim beliefs Π· of the agent if her

prior is Π·?First, if the principal’s ex-ante probability P(p·) = ∑Tt=0 ptΠt of choosing m is

zero, the choice of m is unexpected: the agent’s beliefs are arbitrary. In this case, we will

restrict beliefs to lie in the set ∆Σm defined in (222). On the other hand, if P(p·) is positive,

the agent’s beliefs are uniquely given by Bayes’s Rule: Πt = ptΠt/P(p·) for all t. Hence,

Πt is positive only if pt > 0 which, by (227), implies t /∈ Jm since p· ∈ PΣm. Thus, Πt lies in

∆Σm in this case as well. Combining the two cases, our assumptions imply that for each p·

in PΣm, the agent’s interim beliefs Π· on seeing m will lie in the set

β (p·) =

(

p0Π0

P(p·), ..., pT ΠT

P(p·)

)⊂ ∆Σ

m if P(p·)> 0

∆Σm if P(p·) = 0

(228)

which, in each case, is a subset of ∆Σm. Let β

(PΣ

m)=

β (p·) : p· ∈ PΣm⊂ ∆Σ

m be the image

of PΣm under β .

Claim 40. The correspondence β : PΣm ⇒ ∆Σ

m is upper hemicontinuous and, for any vector

p· ∈ PΣm of probabilities, the set β (p·) is nonempty and convex.

Proof. By (228), β (p·) is nonempty and convex. For upper hemicontinuity, suppose

(p·n,x·n)→ (p·,x·)with p·n ∈ PΣ

m and x·n ∈ β (p·n)for all n. Then p· ∈ PΣm since PΣ

m is compact

(Claim 39). It remains to show that x· ∈ β (p·). Since x·n ∈ β (p·n)⊂ ∆Σm for all n and ∆Σ

m is

closed, x· = limn→∞ x·n is also in ∆Σm; thus, if P(p·) is zero we are done. If P(p·) is positive,

then, since the function P(p·n) is continuous in p·n, there is a n∗ such that, for all n > n∗,

P(p·n) is positive whence x·n ∈ β (p·n) must equal(

p0nΠ0

P(p·n), ...,

pTn ΠT

P(p·n)

)and thus46

x· = limn→∞

x·n = limn→∞

(p0

nΠ0

P(p·n), ...,

pTn ΠT

P(p·n)

)=

(p0Π0

P(p·), ...,

pT ΠT

P(p·)

)∈ β (p·) .

Hence β is upper hemicontinuous.

46This uses the fact that f (x,y) = g(x)/h(y) is continuous at (x0,y0) if g and h are continuous and h(y0) 6=

0.

115

Page 116:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Let Dm denote the Cartesian product Ψm × PΣm × ∆Σ

m, which is compact by Claims

38, 39, and 36, resp. Consider the correspondence r : Dm ⇒ Dm that maps each d =((V ·,U) , p·,Π·

)in Dm to r (d) = ψm

(Π·)×P· (V ·)× β (p·). For any d in Dm, r (d) is

nonempty and convex since ψm

(Π·)

, P· (V ·), and β (p·) are by Claims 37, 39, and 40,

resp. Moreover, r is upper hemicontinuous since ψm, P·, and β have this property by the

same Claims. Thus, by Kakutani’s [9] fixed point theorem, r has a fixed point((V ·∗,U∗) , p·∗,Π

·∗

)∈ Dm (229)

satisfying

(V ·∗,U∗) ∈ ψm

(Π·∗

), (230)

p·∗ ∈ P· (V ·∗) , and (231)

Π·∗ ∈ β (p·∗) . (232)

By (230), there is a correlated equilibrium π∗m ∈ φm

(Π·)

such that

(V ·∗,U∗) =(V ·m (π∗m) ,Um (π∗m)

). (233)

Moreover, the beliefs Π·∗ are in ∆Σm by (229) and thus satisfy IntΣm by Claim 36.

For the mechanism m, we have now constructed beliefs Πt (m) = Π·∗ (which satisfy

IntΣm) and associated correlated equilibrium payoffs(V · (m) ,U (m)

)=(V ·∗,U∗

)of the stage-

2 game. By hypothesis, there is a type t whose payoff V t (m) exceeds V t .

What is the expected type-contingent outcome in the stage-2 equilibrium of m? For

each pure action profile(sP· ,s

A) ∈ (SPm)T+1×SA

m, the probability is π∗m(sP· ,s

A) that(sP· ,s

A)is played. If, moreover, the principal’s type is t, the resulting outcome is osP

t ,sA

m for sure.

Thus, the expected outcome conditional on the principal choosing m and his type being t is

ot∗ = ∑

sA∈SAm

∑sP· ∈(SP

m)T+1

π∗m

(sP· ,s

A)

osPt ,s

A

m . (234)

Now consider the following scenario of the modified game G.

116

Page 117:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Stage 1: the principal of type t proposes m with probability pt∗ and an RSW allocation

o· with complementary probability.

Stage 2: if the principal chose o·, the agent plays ”accept” and the principal chooses

the outcome ot that corresponds to his type t. If the principal chose m, the players play the

correlated equilibrium π∗m whence the expected outcome, for a principal of type t, is ot∗.

We claim that the agent’s ex ante expected payoff in Σ is nonnegative. This payoff is

T

∑t=0

Πt [pt∗U

t (ot∗)+(1− pt

∗)

U t (ot)]≥ T

∑t=0

Πt pt∗U

t (ot∗)

by (140). (235)

If Πt pt∗ is zero for all t, we are done as the right hand side is zero. Else ∑

Ts=0 Πs ps

∗ is

positive, whence the right hand side of (235) has the same sign as

T

∑t=0

Πt pt∗

∑Ts=0 Πs ps

∗U t (ot

∗)=

T

∑t=0

Πt (m)U t (ot

∗)

by (128) with Π· = Π

·

= ∑sA∈SA

m

∑sP· ∈(SP

m)T+1

π∗m

(sP· ,s

A) T

∑t=0

Πt (m)U t

(osP

t ,sA

m

)by (121) and (234)

= ∑sA∈SA

m

∑sP· ∈(SP

m)T+1

π∗m

(sP· ,s

A)

uAm

(sP· ,s

A)

by (127)

= Um (π∗m) by (130)

which is nonnegative by Claim 14.

Now let us give the principal a third option at the initial stage in G: he can propose

the DRM o· defined, for each t, by ot = pt∗o

t∗+(1− pt

∗) ot . If the agent accepts and the

principal chooses ot , the players get the same type-contingent payoffs as in Σ above. Thus,

as Σ gives the agent a nonnegative payoff, she will accept the DRM. And as the principal

does not want to deviate in Σ, he will choose ot : o·is incentive-compatible. But recall

that m is a mechanism for which, for any beliefs Π· ∈ ∆T and any associated equilibrium

payoffs(V ·,U

), there is a type t who does better than in the RSW allocation: V t > V t .

Letting Π· = Π· (m), it follows that there is a type t for which V t∗ > V t and thus, by (230),

pt∗ = 1, so V t (ot) > V t . Since, by (230), V t (ot) = max

V t∗ ,V

t≥ V t , we have produced

an allocation, o·, that is incentive-compatible, Pareto dominates o·, and gives the agent a

117

Page 118:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

nonnegative payoff under the beliefs Π·. Thus, by definition 15, o· is not IE under the

beliefs Π· - a contradiction. Q.E.D.Proposition 3

PROOF OF PROPOSITION 4. Let Σ =(

p··,Π· (·) ,π·

)be an intuitive equilibrium whose

expected allocation is o· = o·Σ= (q·,ρ ·). Assume w.l.o.g. that Σ is a DRM equilibrium.47

When considering deviations by the principal, it will suffice to focus on simple mechanisms

in which the agent has two actions, accept and reject, and the principal has one action: ”do

nothing”. If the agent accepts, a given outcome (q,ρ) is implemented; else the reservation

allocation (in which both get zero) is implemented. We will refer to such a mechanism,

abusing notation slightly, as m = (q,ρ).

We first show that Σ is fairly priced: the amount ρ t that the agent pays to a principal of

each type t equals the conditional expected value qt f (t) of the portfolio that she receives.

For suppose not. First suppose that the portfolio of some type t is underpriced: ρ t < qt f (t).

Let s be the largest type such that qt f (s) < qt f (t). (If there is no such type s, then the

principal can instead select the mechanism m = (qt ,qt f (t)− ε) for any ε ∈ (0,qt f (t)−ρ t)

and the agent will accept so Σ is not an equilibrium.) Choose any λ ∈ [0,1) satisfying

ρ t−δqt f (t)qt [ f (t)−δ f (t)]

< λ <ρ t−δqt f (s)

qt [ f (t)−δ f (s)]. (236)

Such a λ must exist as the ratio on the right is increasing in −qt f (s) since ρ t < qt f (t).

Also choose any

ε ∈(0,λqt [ f (t)−δ f (t)]−ρ

t−δqt f (t))

(237)

where the interval is nonempty by the first inequality in (236). Now suppose type t deviates

to the mechanism m = (λqt ,λqt f (t)− ε). By ASSUMPTION A, for all s′ < s, qt f (s′)≤

qt f (s) < qt f (t) and so, by (236), λqt [ f (t)−δ f (s′)] is less than ρ t − δqt f (s′) which, in

turn, is no greater than V s′(

os′)

since s′ is willing not to imitate t in Σ. Thus, since ε > 0

47If not, use Σ to construct the equivalent DRM equilibrium Σ as in footnote 39. It has the same equilibrium

payoffs of principal and agent for each type of principal. Moreover, beliefs following any deviation m of the

principal are the same in Σ as in Σ. Hence, since πm is intuitive, so is the belief function πm of Σ: We have

thus constructed an intuitive DRM equilibrium that implements the expected allocation o·Σ

of Σ.

118

Page 119:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

by (237), in any stage-2 correlated equilibrium following a deviation to m, each type s′ ≤ s

gets less than her equilibrium payoff V s′(

os′)

: each type s′ ≤ s is in the set Jm defined in

(149). Now suppose the principal chooses m and the agent is sure the agent’s type is not in

Jm. Then by ASSUMPTION A, she is sure that the portfolio λqt is worth at least λqt f (t).

As accepting thus gives her at least ε > 0, she will accept in any correlated equilibrium

πm.48 This implies, in turn, that a type-t principal gains by deviating to m for any πm

since,by (236) and (237), V t (ot) = ρ t −δqt f (t) is less than λqt [ f (t)−δ f (t)]. It follows

from IntΣm that on seeing m, the agent’s interim beliefs Π· (m) must put zero weight on any

type in Jm: she will accept. But then t will surely defect: Σ is not an equilibrium.

We have shown that no portfolio is underpriced in Σ: for each t, ρ t ≤ qt f (t). Can

there be overpricing? No: as each agent has positive ex-ante probability Πt , this (with

the absence of underpricing) would imply that the agent’s ex-ante expected payoff from

accepting the DRM Σ is negative - which contradicts Claim 14.

Since Σ is fairly priced, the price paid by the agent equals the value of the assets she

receives for each type t:

ρt = qt f (t) for t = 0, ...,T. (238)

Hence, the payoff U t (ot) = qt f (t)−ρ t of the agent is identically zero. By Claim 21, the

payoff V t (ot) of the type-t principal is not less than V t which, in turn, equals the solution

u∗ (t) to RLP by Corollary 18. We claim that V t (ot)≤ u∗ (t) as well and that qt lies in the

set Qt defined in (141).If not, let t be the lowest type for which V t (ot) > u∗ (t). By fair

pricing,

V t (ot)= ρt−δqt f (t) equals (1−δ )qt f (t) for each t. (239)

But by (142), (143), and (144), u∗ (t) equals maxq∈Qt [(1−δ )q f (t)]. Thus, if V t (ot) >

u∗ (t), then qt must not lie in Qt . By (141), this means that there is a type s < t whose

equilibrium payoff V s (os) = u∗ (s) in Σ is less than qt [ f (t)−δ f (s)] which in turn, by

48Since the principal’s choice in stage 2 is trivial, correlated equilibrium implies merely that the agent will

act optimally given her beliefs.

119

Page 120:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

(238), equals his payoff ρ t−δqt f (s) from imitating t in Σ. Thus, Σ is not an equilibrium -

a contradiction.

We have shown that in any intuitive equilibrium Σ, the payoff V t (ot) of the type-t

principal equals his RSW payoff V t which, in turn, equals u∗ (t) and the agent’s corresponding

payoff U t (ot) also equals her RSW payoff of zero. Moreover, qt ∈ Qt . If qt does not

maximize qt f (t) in this set then

V t (ot)= (1−δ )qt f (t) by (239)

< (1−δ )q∗ (t) f (t) by (142)

= u∗ (t) by (143) and (144)

=V t (ot) as shown above.

This is a contradiction. We thus have confirmed that q∗ (t) = qt , ρ∗ (t) = ρ t , and u∗ (t) =

V t (ot) are a solution to RLP. Accordingly, by part 2 of Claim 18, the allocation o· is an

RSW allocation as claimed.

We have shown that any DRM o· that satisfies (151), (152), and (153) is supported by

an equilibrium as claimed. Moreover, for each mechanism m ∈M, the property IntΣm holds

since the agent’s interim beliefs Π· (m) have been constructed to lie in ∆Σm. Thus, if o· is

an RSW allocation o·, the equilibrium is intuitive as well: any RSW allocation o· is the

expected allocation of an intuitive equilibrium, as claimed. Q.E.D.Proposition 4

PROOF OF CLAIM 25. Equations (168) through (174) are proved in the text, and (184)

follows from (96), (97), and (170). Substituting (169) into V 2 (o1) = ρ1− δa f (2) we

obtain the equality in (175). The RSW allocation o· is an equilibrium allocation by Claim

24 and q2 6= q1 = a by Claim 25, whence the inequality in (175) follows from the following

lemma.

Lemma 8. In any equilibrium allocation o· in which q2 6= q1 = a, V 2 (o2)>V 2 (o1).

120

Page 121:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Proof. By type 1’s IC constraint,

V 2 (o2)−V 2 (o1)≥V 2 (o2)−V 2 (o1)− [V 1 (o2)−V 1 (o1)]= δ

(a−q2)

∆1 > 0

where the strict inequality is by (154).

By (170) and (174), V 1− V 2 equals q2 [∆− (1−δ ) f (2)] which in turn, by (97), equals

δ q2∆1, proving (176). As for (178),

V 1−V 2 (o1)= (1−δ )a f (1)−a [ f (1)−δ f (2)] = δa∆1

as claimed whence, by (176),

V 2−V 2 (o1)= [V 1−V 2 (o1)]−[V 1−V 2]= δ

(a− q2)

∆1

proving (179). Equation (180) follows from (164) and (165). As for (181) and (182), there

are two cases.

1. If V 1 ≤ ∆1, (171) and (172) imply q21 =

V 1

∆1and q2

2 = 0. Substituting these into (174),

clearing denominators, and multiplying by−1, we obtain (δ −1)V 1 f1 (2) =−V 2∆1.

Adding ∆1V 1 to both sides then yields δV 1∆11 = ∆1

(V 1−V 2

), which proves (a) the

< and = parts of (183), (b) that the second elements of the mins in (171) and (181)

are equal, and (c) that the second entry of the max in (181) is zero. This verifies (181)

and (182) for this case.

2. If V 1 > ∆1, (171) and (172) imply q21 = 1 and q2

2 = V 1−∆1∆2

. Substituting these into

(174) and clearing denominators yields

V 2∆2 = (1−δ )

[f1 (2)∆2 +

(V 1−∆1

)f2 (2)

].

Isolating the ∆2 terms on the right hand side yields(V 1−∆1

)[(δ −1) f2 (2)] = ∆2

−V 2− (δ −1) f1 (2)

.

121

Page 122:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Adding(

V 1−∆1

)∆2 to both sides and using (97) to substitute for ∆2 on the left and

for ∆1 on the right we obtain(V 1−∆1

)δ∆

12 = ∆2

V 1−V 2−δ∆

11

,

which (a) proves the > part of (183) and (b) implies that the second elements in the

maxes in (172) and (182) are equal. Hence, by hypothesis that V 1 > ∆1, the second

element of the max in (182) is positive, whence the second element in the min in

(181) exceeds 1. Finally, This verifies (181) and (182) for this case.

As for (177), by (171), (172), and (176), there are two cases.

1. If V 1 ≤ ∆1 then q2 =(

V 1

∆1,0)

so

V 1−V 2 = δV 1

∆1∆

11. (240)

2. If V 1 > ∆1 then q2 =(

1, V 1−∆1∆2

)so

V 1−V 2 = δ

11 +

V 1−∆1

∆2∆

12

. (241)

Combining the two cases we obtain (177). Finally, we verify (184). We compute

ζ =γ

a∆1 >δ∆1

1∆1

= η ⇐⇒ ∆11a∆

1 <∆1γ

δ= ∆1

(a− q2)

∆1 = ∆1a∆

1−∆1q2∆

1

⇐⇒0 <(∆1−∆

11)

a∆1−∆1q2

∆11 (242)

When ν > 0, by (171), (172), and (183), this becomes

0 <(∆1−∆

11)

a∆1−∆1

V 1

∆1∆

11 = (1−δ ) f1 (1)a∆

1−V 1∆

11

= (1−δ ) f1 (1)a [ f (2)− f (1)]− (1−δ )a f (1) [ f1 (2)− f1 (1)]

= (1−δ ) f1 (1)a [ f (2)− f (1)]−a f (1) [ f1 (2)− f1 (1)]

= (1−δ ) f1 (1) [ f1 (2)− f1 (1)]+ f1 (1) [ f2 (2)− f2 (1)]− [ f1 (1)+ f2 (1)] [ f1 (2)− f1 (1)]

= (1−δ ) f1 (1) [ f2 (2)− f2 (1)]− f2 (1) [ f1 (2)− f1 (1)]

= (1−δ ) f1 (1) f2 (2)− f2 (1) f1 (2)

122

Page 123:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

which is true by (155). When ν ≤ 0, by (171), (172), and (183), (242) becomes

0 <(∆1−∆

11)

a∆1−∆1

(∆

11 +

V 1−∆1

∆2∆

12

)

=−∆11a∆

1 +∆1

(a∆

1−∆11−

V 1−∆1

∆2∆

12

)

=−∆11a∆

1 +∆1∆1

2∆2

(a∆−V 1

)=−∆

11a∆

1 +∆1∆1

2∆2

a∆1 by (184)

= ∆1

[∆1

2∆2−

∆11

∆1

]a∆

1

which is positive by (154) and (162).

Q.E.D.Claim 25

PROOF OF CLAIM 26. By Claim 25, it suffices to show that o· is an RSW allocation.

Since V 1 = (1−δ )a f (1) is the maximum gains from trade when the seller’s type is 1,

V 1 (o1)+U1 (o1) ≤ V 1 = V 1 (o1) whence U1 (o1) ≤ 0 and thus, by (158), U2 (o2) ≥0. First suppose U1 (o1) = 0. Then o· clearly satisfies (140). As o· is an equilibrium

allocation, it also satisfies (139). Since, moreover, it gives each type of seller her RSW

payoff, it solves Program IIt in Definition 14 and hence is an RSW allocation as claimed.

Now suppose instead that U1 (o1)< 0, whence U2 (o2)> 0; we will prove a contradiction.

Consider the alternative allocation o· defined by o1 = o1 and o2 = o2. We have

U1 (o1)=U1 (o1)= 0 by part 1 of Claim 18;

U2 (o2)=U2 (o2)> 0 by assumption;

V 1 (o2)=V 1 (o2)≤V 1 (o1)=V 1 (o1)=V 1 (o1) by construction, as o· is an equilibrium,

by hypothesis, and by construction, resp.;

V 2 (o2)=V 2 (o2)=V 2 (o2)≥V 2 (o1)=V 2 (o1) by construction, by hypothesis,

by Claim 17, and by construction, resp.

123

Page 124:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Accordingly, o· satisfies (139) and (140) and gives each type of seller her RSW payoff.

Thus,o· solves Program IIt in Definition 14 and so is an RSW allocation. However, U2 (o2)>0, which contradicts part 1 of Claim 18.

Q.E.D.Claim 26

PROOF OF CLAIM 27. By part (c) of Proposition 3, the payoff V 1 (o1) of the low type

seller cannot be less than her RSW payoff V 1 which, by parts 1 of Claims 10 and Claim

18, equals (1−δ )a f (1). But this is the maximum gains from trade V 1 (o1)+U1 (o1) =(1−δ )q1 f (1) when the seller’s type is low. Hence U1 (o1) ≤ 0 and thus, by (158),

U2 (o2) ≥ 0. Accordingly, if U1 (o1) = 0, then o· satisfies (140). As o· is an equilibrium

allocation, it also satisfies (139) and thus, by Definition 14, V t (ot) ≤ V t for each t, which

contradicts P0. We conclude that U1 (o1)< 0 whence, by (158), U2 (o2)> 0.

Now suppose that V 1 (o1)≤ V 1; we will derive a contradiction. By (159), V 1 (o1)= V 1,

and by P0, V 2 (o2)> V 2. By part 2 of Claim 18, the above unique solution to RLP induces

an RSW allocation o· defined by ot = (qt , ρ t) = (q∗ (t) ,ρ∗ (t)) for t = 1,2. Consider the

allocation o· defined by o1 = o1 and o2 = o2. We have

U1 (o1)=U1 (o1)= 0 by part 1 of Claim 18;

U2 (o2)=U2 (o2)> 0 as shown above;

V 1 (o2)=V 1 (o2)≤V 1 (o1)=V 1 (o1)=V 1 (o1) by construction, as o· is an equilibrium,

as shown above, and by construction, resp.;

V 2 (o2)=V 2 (o2)≥V 2 (o2)≥V 2 (o1)=V 2 (o1) by construction, as shown above,

by Claim 17, and by construction, resp.

Accordingly, o· satisfies (139) and (140) yet gives type 2 seller more than her RSW payoff

- contradicting Definition 14.

Q.E.D.Claim 27

PROOF OF CLAIM 28. The set S is compact by Claim 24. Hence, Sx is also compact

124

Page 125:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

as it adds an equality to the set of constraints that define S. Thus, if Sx is nonempty, then

the continuous function J has a maximum on Sx by the Extreme Value Theorem: Sx is

nonempty as well.

Now suppose Sx is nonempty and let o· ∈ Sx. We will show that o· satisfies P1-P4 in turn.

Suppose first that o· violates P1. First suppose q1 6= a. Let q1 = a and ρ1 = ρ1+δ(a−q1) f (1),

and o2 = o2. One can see that V t (ot) =V t (ot) for t = 1,2: o· satisfies P0 and (159) since

o· does. Moreover, V 1 (o2)=V 1 (o2) since o2 = o2, and

V 2 (o1)−V 2 (o1)= ρ1−ρ

1−δ(a−q1) f (2) = δ

(a−q1) [ f (1)− f (2)]≤ 0

whence o· satisfies (156) since o· does. Finally,

U1 (o1)−U1 (o1)= (1−δ )(a−q1) f (1)> 0 (243)

by (154) while U2 (o2) = U2 (o2) whence o· satisfies (158) as o· does: o· is in Sx. But

by (243), J (o·) > J (o·): o· is not in Sx, a contradiction. Finally, using q1 = a, we set

V 1 (o1) = ρ1− δa f (1) equal to the required payoff of V 1 + x and use (170) to obtain

ρ1 = a f (1)+ x.

Now say o· violates P2. By (156), this implies V 1 (o2)<V 1 (o1). Hence q2 6= a since

otherwise type 2 would deviate to o1.49 For any ε > 0, consider the alternative allocation o·

defined by o1 = o1, q2 = (1− ε)q2+εa, and ρ2 = ρ2+δ(q2−q2) f (2). By construction,

V 1 (o1)=V 1 (o1)= V 1+x, V 2 (o2)=V 2 (o2), and U2 (o2)>U2 (o2), which imply (157),

(160), and (158). Finally, since V 1 (o2)<V 1 (o1), we can choose ε > 0 small enough that

(156) holds so o· is in Sx whence (as U2 (o2)>U2 (o2)) the original allocation o· is not in

Sx - a contradiction.

Suppose now that o· violates P3: q2 6= a and V 2 (o2) > V 2. For any ε > 0, consider

the alternative allocation o· defined by o1 = o1, q2 = (1− ε)q2 + εa, and ρ2 = ρ2 +

δ(q2−q2) f (1). By construction, V 1 (o1)=V 1 (o1)= V 1+x. By P2, V 1 (o2)=V 1 (o2)=

V 1 (o1)=V 1 (o1) and U2 (o2)>U2 (o2)which imply (156) and (158), respectively. Moreover,

49If q2 = a = q1 then V 2(o1)−V 2

(o2)= ρ1−ρ2 =V 1

(o1)−V 1

(o2)> 0.

125

Page 126:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

by P1 and P2,

V 2 (o2)−V 2 (o1)= ρ2−ρ

1 +δ(a−q2) f (2)

=V 1 (o2)−V 1 (o1)+δ(a−q2)

∆1

>V 1 (o2)−V 1 (o1)= 0

where the inequality relies on ASSUMPTION A and genericity. Thus, if ε is small enough,

(157) and (160) hold whence o· is in Sx. But since U2 (o2)−U2 (o2)> 0, o· is not in Sx -

a contradiction.

Finally, suppose o· violates P4: q21 < 1, q2

2 > 0, and hence, by P3, V 2 (o2)= V 2. Let us

now raise q21 slightly while lowering q2

2 and adjusting ρ2 so as to leave the payoffs of each

type of seller from o2 unchanged. We will show that this makes the agent better off when

facing the type-2 seller, so o· is not in Sx - a contradiction. For any small constants ε, ι ,ω ,

consider the alternative allocation o· defined by o1 = o1 = o1, q21 = q2

1+ε , q22 = q2

2− ι , and

ρ2 = ρ2+ω . Given ε > 0, we seek ι and ω such that the change does not alter the payoff of

either seller from choosing the outcome designated for either seller, whence (156), (157),

(159), and (160) still hold. As o1 = o1, V t (o1)=V t (o1) for t = 1,2. Now,

V 1 (o2)−V 1 (o2)= ω−δε f1 (1)+δι f2 (1)

and

V 2 (o2)−V 2 (o2)= ω−δε f1 (2)+δι f2 (2) .

Subtracting and dividing by δ , the two expressions are equal as long as ι = εψ where

ψ = ∆11/∆1

2 is positive by (154). Finally, the second expression (and thus the first) is zero

as long as ω = δε [ f1 (2)−ψ f2 (2)]. With these choices, the agent’s payoff from selling to

type 2 changes by

U2 (o2) − U2 (o2) = ε f1 (2) − ι f2 (2) − ω = ε (1−δ ) [ f1 (2)−ψ f2 (2)]

If we multiply the term in square brackets by ∆12, which is positive by (154), it becomes

∆12 f1 (2) − ∆

11 f2 (2) = f1 (1) f2 (1)

[f2 (2)f2 (1)

− f1 (2)f1 (1)

]

126

Page 127:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

which is positive by (155). Accordingly, U2 (o2)>U2 (o2), so o· is not in Sx - a contradiction.

We conclude that o· satisfies P1, P2, P3, and P4, as claimed. Q.E.D.Claim 28

PROOF OF CLAIM 29. By Claim 28, any such allocation satisfies P1-P4. Given P1,

q1 = q2 implies q2 = a. Moreover, by P2, ρ2 = ρ1 = a f (1)+x: a pooling menu is offered.

As the agent’s payoff must be nonnegative, using the notation U (o·) of (158),

0≤U (o·) =2

∑t=1

Πt (qt f (t)−ρ

t)= a2

∑t=1

Πt f (t)− [a f (1)+ x]

= Π2a∆

1− x.

Accordingly, we must have x ≤ Π2a∆1. By Claim 24, for this to be an equilibrium it

suffices that the type 2 seller gets at least his RSW payoff: that V 2 (o2) = ρ1−δa f (2) =

a [ f (1)−δ f (2)]+x is not less than V 2 or, equivalently, x≥ V 2−a [ f (1)−δ f (2)] = V 2−

V 2 (o1). Combining these inequalities and using (179), x must lies in the interval (194) as

claimed. Moreover, for this interval to be nonempty, we must have δ(a− q2)∆1 ≤Π2a∆1,

which is equivalent to (193) as claimed.50 Q.E.D.Claim 29

PROOF OF PROPOSITION 5. Let o· ∈ Sx satisfy q2 6= q2. By P2 we have V 1 (o2) =ρ2−δq2 f (1) =V 1 (o1)= V 1 + x, whence

ρ2 = δq2 f (1)+V 1 + x. (244)

Combining (195) and (244) to eliminate ρ2 yields

x = δq2∆

1−(

V 1−V 2). (245)

By P4, there are now two subcases.

1. Assume (186). Then by (171), (181), and (182), type 2’s RSW outcome satisfies

(188). There are now two intervals in which x may lie.

50The right hand side of (193) is positive by (154) and (174). It is less than one since q2 6= a by (171) and

(172) and thus V 2 = (1−δ ) q2 f (2) (using (174)) is less than (1−δ )a f (2) =V 2(o1)+a∆1 by (154).

127

Page 128:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

(a) If x lies in (196) then, by (154) and (245), q2 is given by (198) so, as x→ 0, q2

converges to q2 by (188) and hence, by (195), ρ2 converges to δ q2 f (2)+ V 2

which equals ρ2 by (173) and (174). It follows that the agent’s expected payoff

(158) converges to her RSW payoff of zero as x→ 0. For x > 0, her type-

contingent payoff U t (ot) = qt f (t)−ρ t equals−x when t = 1 and, using (195),

U2 (o2)= q2 f (2)−ρ2 = (1−δ )q2 f (2)−V 2

= (1−δ )V 1−V 2 + x

δ∆11

f1 (2)−V 2

when t = 2. Hence, using Π1 = 1−Π2, her unconditional expected payoff

(158), for a given increment x in type 1’s payoff, is

U (o·) =U (o·|x) =−x+Π2

[(1−δ )

V 1−V 2 + xδ∆1

1f1 (2)+ x−V 2

](246)

which at the top endpoint of (196) equalsV 1−V 2−δ∆11 +Π2

[∆1−V 1

]which

by (187) is nonnegative if and only if51

Π2 ≥

δ∆11−(

V 1−V 2)

∆1−V 1= η , (247)

which confirms (197). Alternatively, one can differentiate (246) with respect to

x:

ddx

U (o·|x) =−1+Π2 (1−δ ) f1 (2)+δ∆1

1δ∆1

1=−1+Π

2 ∆1

δ∆11=−1+

Π2

η.

(248)

As U (o·|x) is zero when x = 0, the constraint (158) holds at x if and only if this

derivative is nonnegative, which is equivalent to (197).

51By (97), (181), (181), and (186),

δ∆11−(

V 1−V 2)

∆1−V 1=

δ∆11

∆1

1− V 1−V 2

δ∆11

1−V 1/∆1=

δ∆11

∆1= η .

128

Page 129:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

(b) If

x≥ ν (249)

then, by (245), q2 is given by (201). The requirement that q22 ≤ 1 then imposes

the additional constraint on x that

x≤ δa∆1−(

V 1−V 2)= γ (250)

by (179) and (164); thus, x must lie in (199) as claimed.52 It remains to check

that the agent’s payoff (158) is nonnegative. As in part 1(a), her type-contingent

payoff U t (ot) = qt f (t)−ρ t satisfies

U1 (o1)=−x and (251)

U2 (o2)= (1−δ )q2 f (2)−V 2. (252)

Thus, using (201) and Π1 = 1−Π2, condition (158) can be rewritten as

0≤U (o·|x) =−x+Π2

(1−δ )

[f1 (2)+ f2 (2)

V 1−V 2 + x−δ∆11

δ∆12

]+ x−V 2

.

(253)

The left hand side is linear and continuous in x, so U (o·|x)=U (o·|ν)+(x−ν)∂U (o·|x)/∂x

as ∂U (o·|x)/∂x is constant over x > ν . Thus, using (253) and (165) to compute

U (o·|ν) =−ν +Π2[(1−δ ) f1 (2)+ν−V 2

]=−ν +Π

2(

∆1−V 1)

and using (253) to compute

∂U (o·|x)∂x

=−x+Π2(1−δ ) f2 (2)

1δ∆1

2+1=−1+Π

2 ∆2

δ∆12,

we find that U (o·|x) is nonnegative at x in (199) if and only if

0≤U (o·|x) =−ν +Π2(

∆1−V 1)+(x−ν)

[−1+Π

2 ∆2

δ∆12

]=−x+Π

2[

∆1−V 1 +(x−ν)∆2

δ∆12

](254)

52Intuitively, for o· to be an equilibrium, type 2’s payoff V 2(o1)+ x from type 1’s outcome o1 cannot

exceed her payoff V 2(o2)

from her own outcome which, in turn, equals V 2 by P3.

129

Page 130:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

or equivalently if Π2 ≥ φ (x) (defined in (200)) as claimed. We next show that

φ (x) is in (0,1) for x in (199). As x≥ ν > 0, it suffices to show that

x < ∆1−V 1 +(x−ν)∆2

δ∆12

⇐⇒x[

1− ∆2

δ∆12

]< ∆1−V 1−ν

∆2

δ∆12

⇐⇒− (1−δ ) f2 (2)δ∆1

2x < ∆1−V 1−ν

∆2

δ∆12

⇐⇒0 < (1−δ ) f2 (2)x+(

∆1−V 1)

δ∆12−∆2ν

As this expression is increasing in x, it suffices to check the inequality at x = ν

when it becomes

0 <(

∆1−V 1−ν

)δ∆

12

or, equivalently, by (154),

0 < ∆1−V 1−ν = ∆1−V 1−δ∆11 +(

V 1−V 2)

= (1−δ ) f1 (2)−V 2 = (1−δ )[

f1 (2)− q2 f (2)]

by (174)

= (1−δ )[∆1−V 1

] f1 (2)∆1

by (171), (172), and (187)

which is positive by (154) and (187) as claimed. Next, φ (x) is of the form xα+βx

where β = ∆2δ∆1

2> 1 by (163), so

φ′ (x) =

α

(α +βx)2

which is clearly finite (so φ is continuous). It is positive if and only if

0 < α = ∆1−V 1−ν∆2

δ∆12

⇐⇒δ∆12

(∆1−V 1

)> ν∆2 =

[δ∆

11−(

V 1−V 2)]

∆2 =

[δ∆

11−δ

V 1

∆1∆

11

]∆2 by (188)

⇐⇒δ∆12

(∆1−V 1

)∆1 > δ∆

11

(∆1−V 1

)∆2⇐⇒ ∆

12∆1 > ∆

11∆2

130

Page 131:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

which holds by (162). We next must evaluate φ (x) at the endpoints of (199).

At the lower endpoint we obtain

φ (ν) =ν

∆1−V 1=

δ∆11−δ q2∆1

∆1−V 1by (180) and (176)

=δ∆1

1−δV 1

∆1∆1

1

∆1−V 1=

δ∆11

∆1= η

as claimed, while at the upper endpoint we obtain

φ (γ) =γ

∆1−V 1 +(γ−ν) ∆2δ∆1

2

∆1−V 1 +δ∆12

∆2δ∆1

2

∆1 +∆2−V 1=

γ

a∆− (1−δ )a f (1)=

γ

a∆1 = ζ by (175) and (179)

as claimed. Next, it is trivial from (200) to see that limx→∞ φ (x) equals δ∆12/∆2;

this exceeds ζ as φ is increasing and φ (γ) = ζ . Finally, since φ is continuous

and increasing on [ν ,γ] and maps this interval to [η ,ζ ], it has a continuous

and increasing inverse function φ−1 defined on [η ,ζ ] that maps this interval to

[ν ,γ], and for(x,Π2) ∈ [ν ,γ]× [η ,ζ ] , Π2 ≥ φ (x) if and only if x≤ φ−1 (Π2).

Hence, the region R defined by (199) and (200) is the union of two regions. The

first, region (i), is the subset of R on which Π2 is in [η ,ζ ] . Since, at all points

in R, x lies in [ν ,γ] , Π2 ≥ φ (x) is equivalent to x ≤ φ−1 (Π2). Hence, region

(i) is defined by the two conditions (202) and (203). The second, region (ii),

is the subset of R on which (205) holds. Since φ (x) ≤ ζ when x is in [ν ,γ],

condition (200) holds automatically if, in addition, (199) holds, region (ii) is

defined simply by (205) and (199).

2. Suppose (189) holds. In this case, x> 0 clearly cannot be less than δ∆11−(

V 1−V 2)

.

Hence, reasoning as in case (a)(ii), x must lie in (206) and q2 must be given by (201).

This has two implications. First, since q22≤ 1, x must lie in (206) as claimed. Second,

as x→ 0, q2 converges to q2 by (191) and hence, by (195), ρ2 converges to δ q2 f (2)+

V 2 which equals ρ2 by (173) and (174). So as in case (a)(i), the agent’s expected

payoff (158) converges to her RSW payoff of zero as x→ 0. Since, moreover, the

131

Page 132:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

agent’s expected payoff is again given by (253), a necessary and sufficient condition

for there to be a separating equilibrium with x in (206) is that

0≤ ddx

U (o·|x) =−1+Π2(1−δ ) f2 (2)

δ∆12

+1=−1+Π

2 ∆2

δ∆12

or, equivalently, that

Π2 ≥

δ∆12

∆2. (255)

Since (189),

γ∆2

δ= ∆2

(a− q2)

∆1 by (164)

= ∆2

((1,1)−

(1,

V 1−∆1

∆2

))∆

1 by (172) and (181)

= ∆2

(1− V 1−∆1

∆2

)∆

12 =

(∆1 +∆2−V 1

)∆

12 = ∆

12 a [ f (2)−δ f (1)]− (1−δ )a f (1)

= ∆12a∆

1

and thus δ∆12

∆2= γ

a∆1 , whence, (255) is equivalent to (207) when (189) holds. Q.E.D.Proposition 5

PROOF OF CLAIM 30. Clearly, if o· is in S, it must also be in Sx for that x equal to the

(constant) increment V 1 (o1)− V 1 that the type 1 seller gets in o·. Thus, to characterize

S, we can restrict attention to allocations that are in Sx for some payoff increment x. Any

element of S must therefore satisfy P1 and P4: type 1 sells her whole portfolio and type

2 sells all of asset 1 before she sells any of asset 2. Moreover, by (192), any element of

S must maximize (in S) the value ∑2t=1 Πtqt f (t) of assets sold. Hence, every allocation in

Spoolx (for any x) must be in S since, in each such allocation, the type 2 seller sells her entire

portfolio. Moreover, if Spoolx is nonempty for some x, then no element of any Ssep

x can be in

S as any such allocation involves type 2 retaining some units of asset 2. Finally, by Claim

29, Spoolx is nonempty for some x if and only if Spool

γ is nonempty, which is so if Π2 ≥ ζ . It

follows that, if Π2 ≥ ζ , S consists only of pooling allocations in which, for some x in the

interval (194), the agent pays ρ t = a f (1)+x to each type of seller for her whole portfolio.

132

Page 133:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

Now suppose Π2 < γ: Spoolx is empty for any x > 0. Hence, any element of S must be in

Ssepx for some x > 0. But by Proposition 5, an increase in x always leads to higher welfare

as it leads type 2 to sell more of one asset and no less of the other. Thus, for any Π2 < γ ,

S consists of the (by Proposition 5) singleton element of Ssepx for the highest x for which

Ssepx is nonempty. As there is no such x when either (a) ν > 0 and Π2 < η or (b) ν ≤ 0 and

Π2 < γ , S is empty in these cases. But if ν > 0 and Π2 ∈ [η ,γ), the highest x for which

Ssepx is nonempty is x = φ−1 (Π2) by case 1(b) of Proposition 5. This proves the result.

Q.E.D.Claim 30

References

[1] Apostol, T. M. (1981). Mathematical Analysis (2nd ed.), Reading, Mass.: Addison-

Wesley.

[2] Aumann. Robert. 1974. ”Subjectivity and Correlation in Randomized Strategies.”

Journal of Mathematical Economics 1:67-96.

[3] Bertsekas, Dimitri P. 2009. Convex Optimization Theory. Athena Scientific.

[4] Cho, In-Koo, and David M. Kreps. 1987. ”Signaling Games and Stable Equilibria.”

Quarterly Journal of Economics 102:179-221.

[5] Cotter, Kevin D. 1991. ”Correlated Equilibrium in Games with Type-Dependent

Strategies.” Journal of Economic Theory 54:48-68.

[6] Forges, Francoise. 2006. “Correlated Equilibrium in Games with Incomplete

Information Revisited.” Theory and Decision 61:329-344.

[7] Fudenberg, Drew, and Jean Tirole. 1991. ”Perfect Bayesian Equilibrium and

Sequential Equilibrium.” Journal of Economic Theory 53:236-260.

[8] Hart, Sergiu, and David Schmeidler. 1989. ”Existence of Correlated Equilibria.”

Mathematics of Operations Research 14:18-25.

133

Page 134:  · 1 Online Appendix to “Portfolio Liquidity and Security Design with Private Information”† Peter M. DeMarzo (Stanford University) David M. Frankel (Melbourne Business School

[9] Kakutani, Shizuo. 1941. ”A generalization of Brouwer’s fixed point theorem.” Duke

Mathematical Journal 8:457–459.

[10] Mailath, G., and E.-L. von Thadden. (2013). Incentive compatibility and

differentiability: new results and classic applications, Journal of Economic Theory

148:1841–1861.

[11] Mas-Collel, Andreu, Michael D. Whinston, and Jerry R. Greene. 1995.

Microeconomic Theory. Oxford:Oxford University Press.

[12] Maskin, Eric, and Jean Tirole. 1992. ”The Principal-Agent Relationship with an

Informed Principal, II:Common Values.” Econometrica 60:1-42.

[13] Myerson, Roger B. 1982. ”Optimal Coordination Mechanisms in Generalized

Principal-Agent Games.” Journal of Mathematical Economics 10:67-81.

[14] Myerson, Roger B. 1983. ”Mechanism Design by an Informed Principal.”

Econometrica 51:1767-1797.

[15] Nash, John F. 1950. ”Equilibrium points in n-Person Games.” Proceedings of the

National Academy of Sciences of the United States of America 36:48-49.

[16] Pontryagin, L.S., V.G. Boltyanskii, R.V. Gamkrelidze and E.F. Mishchenko (1962),

The mathematical theory of optimal processes, New York: Interscience.

134


Recommended