dynamic discrete choice under rational inattentionpeople.bu.edu/miaoj/ups04.pdfma ckowiak and...

45
Dynamic Discrete Choice under Rational Inattention * Jianjun Miao Hao Xing May 12, 2020 Abstract We adopt the posterior-based approach to study dynamic discrete choice problems under rational inattention. We provide necessary and sufficient conditions to characterize the solution for the additive class of uniformly posterior-separable cost functions. We propose an efficient algorithm to solve these conditions and apply our model to explain phenomena such as status quo bias, confirmation bias, and belief polarization. A key condition for our approach to work is the concavity of the difference between the generalized entropy of the current posterior and the discounted generalized entropy of the prior beliefs about the future states. Keywords: Rational Inattention, Endogenous Information Acquisition, Entropy, Dynamic Dis- crete Choice, Dynamic Programming JEL Classifications: D11, D81, D83. * We thank John Leahy, Bart Lipman, Stephen Morris, Jakub Steiner, Colin Stewart, and Mike Woodford for helpful discussions. Department of Economics, Boston University, 270 Bay State Road, Boston MA 02215. Email: [email protected]. Tel: (617) 3536675 Department of Finance, Questrom School of Business, Boston University, 595 Commonwealth Ave, Boston MA 02215. Email: [email protected].

Upload: others

Post on 22-Jan-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Dynamic Discrete Choice under Rational Inattention ∗

Jianjun Miao† Hao Xing‡

May 12, 2020

Abstract

We adopt the posterior-based approach to study dynamic discrete choice problems underrational inattention. We provide necessary and sufficient conditions to characterize the solutionfor the additive class of uniformly posterior-separable cost functions. We propose an efficientalgorithm to solve these conditions and apply our model to explain phenomena such as statusquo bias, confirmation bias, and belief polarization. A key condition for our approach to workis the concavity of the difference between the generalized entropy of the current posterior andthe discounted generalized entropy of the prior beliefs about the future states.

Keywords: Rational Inattention, Endogenous Information Acquisition, Entropy, Dynamic Dis-crete Choice, Dynamic Programming

JEL Classifications: D11, D81, D83.

∗We thank John Leahy, Bart Lipman, Stephen Morris, Jakub Steiner, Colin Stewart, and Mike Woodford forhelpful discussions.†Department of Economics, Boston University, 270 Bay State Road, Boston MA 02215. Email: [email protected].

Tel: (617) 3536675‡Department of Finance, Questrom School of Business, Boston University, 595 Commonwealth Ave, Boston MA

02215. Email: [email protected].

Page 2: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

1 Introduction

Economic agents often make dynamic discrete choices, such as whether to stay at home or take

a job and which job to take, when to replace a car and which new car to buy, when to invest in

a project and which project to invest, and so on. When making these decisions people often face

imperfect information about payoffs. People must choose what information to acquire and when

to acquire it given their limited attention to the available information.

We adopt the rational inattention (RI) framework introduced by Sims (1998, 2003) to study

the optimal information acquisition and choice behavior in a dynamic discrete choice model. In

the model a decision maker (DM) can choose a signal about a payoff-relevant state of the world

before taking an action in each period. The state follows a finite Markov chain with a transition

kernel depending on the current states and actions. The DM receives flow utilities, that depend

on the current states and chosen actions, and pays a utility cost to acquire information, that is

proportional to the reduction in the uncertainty measured by a generalized entropy function of his

beliefs. The DM’s objective is to maximize the expected discounted utility less the cost of the

information he acquires. We call this problem the dynamic RI problem.

The existing literature typically adopts the Shannon (1948) entropy cost function. Despite many

appealing features of this specification, the experimental literature in economics and psychology

establishes some behavior that violates key features of the Shannon model (see, e.g., Woodford

(2012), Caplin and Dean (2013) (henceforth CD), and Dewan and Neligh (2020)). Motivated by

this evidence, CD (2013) and Caplin, Dean, and Leahy (2019b) (henceforth CDL)) propose more

flexible cost functions. While they provide solutions in a static setup given these cost functions,

how to extend their analysis to a dynamic setup is still an open question. The goal of our paper is

to fill this gap.

We make three contributions to the literature. First, we characterize the solution to the dynamic

RI problem using the posterior-based approach. To apply this approach, we focus on the class of

uniformly posterior-separable (UPS) cost functions proposed by CD (2013) and CDL (2019b).

Solving the dynamic RI problem is difficult because the current information acquisition affects

future beliefs, which in turn influence the continuation value in a nonlinear way. The continuation

value may not be concave in the revised prior beliefs following any history reached with positive

probabilities.1 By dynamic programming, the current choice and the continuation value are linked

by the Bellman equation. It is unclear whether this dynamic programming problem is concave.

Steiner, Stewart, and Matejka (2017) (henceforth SSM) solve the dynamic RI problem in the

special case of Shannon entropy using the choice-based approach. They first transform the problem

into an unconstrained control problem and then take coordinate-wise first-order conditions to pro-

1We can show that it is actually convex for the Shannon entropy case.

2

Page 3: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

vide a dynamic logit characterization (Rust (1987)). We argue that this approach does not work

for general UPS cost functions. Our posterior-based approach is built on the insights of CD (2013)

in a static model and takes into account the issue of joint concavity in a dynamic setting. We

derive the posterior-based Bellman equation using the predictive distribution as the state variable.

This distribution given any history can be viewed as the prior belief about the future states at that

history. It is revised from the current posterior through the state transition kernel.

We reduce the dynamic RI problem to a collection of static problems using the Bellman equation.

The static problem in each period is to solve the concavification of a collection of net utilities as

functions of the posterior. Each net utility function consists of the current net utility and the

continuation value. It is critical for this function to be concave for the concavification problem to

be solvable. We show that the overall net utility is concave under the assumption that the difference

between the generalized entropy of the current posterior and the discounted generalized entropy

of the prior belief about the future states is concave. This assumption is also important for us to

establish a recommendation lemma similar to those in SSM (2017) and Ravid (2019), which states

that the signal-based formulation with imperfect information is equivalent to our posterior-based

formulation with full information.

When further restricting to the class of generalized entropy functions that are additive across

states, we provide a tractable first-order characterization for the dynamic RI problem using the

result in CD (2013). This characterization gives necessary and sufficient conditions for optimal

solutions. It reduces to the dynamic logit characterization of SSM (2017) in the special Shannon

entropy case.

Our second contribution is to propose a characterization of Markovian solutions and an efficient

algorithm to find such a solution. For a Markovian solution, the predictive distribution of the

next-period states depends only on the current action, the default rule depends only on the last

period action, and the choice rule depends only on the current state and the last period action.

Our characterization generalizes that of SSM (2017) by allowing corner solutions and UPS cost

functions.

Our algorithm extends the forward-backward Arimoto-Blahut algorithm of Tanaka, Sandberg,

and Skoglund (2018) to infinite-horizon models with discounting and to generalized entropy func-

tions. This algorithm is based on the Arimoto-Blahut algorithm for solving static channel capacity

and rate distortion problems with Shannon entropy in information theory in the engineering liter-

ature (Arimoto (1972) and Blahut (1972)).

Our third contribution is to apply our theoretical results and numerical methods to solve some

economic examples based on a matching state problem often studied in the literature (e.g., CD

(2013), SSM (2017), and CDL (2019a)). We show that RI can help explain some phenomena

documented in the psychology literature, such as status quo bias, confirmation bias, and belief

3

Page 4: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

polarization. We find that the status quo bias discussed by SSM does not arise when the decision

horizon is sufficiently long. The reason is that the probability of switching states in the future is

getting larger if the horizon is longer. Thus the DM has incentive to acquire new information and

take a different action. We also show that there is a positive feedback between beliefs and actions

when the state transition kernel depends on actions. This property is useful to understand the

preceding behavioral biases.

Our numerical examples adopt the Cressie and Read (1984) entropy function that includes

Shannon entropy as a special case. The Cressi-Read entropy function incorporates a curvature

parameter that affects the marginal information cost, thereby affecting both static and dynamic

choice probabilities as well as the timing of choices. We find that the status quo bias can occur

earlier and the confirmation bias is more likely to occur when the curvature parameter is smaller

because it induces a larger marginal information cost.

Our paper is closely related to CD (2013, 2015), Matejka and McKay (2015) (henceforth MM),

SSM (2017), and CDL (2019a, b). SSM (2017) is the first paper that extends the static model of

MM (2015) to a dynamic setting and derives the dynamic logit rule.2 Their solution method does

not apply to the general UPS cost functions adopted in our paper. We also extend the SSM model

to allow the state transition kernel to depend on actions. Our generalization permits us to study a

wide range of economic and psychological behavior.

Our paper is also related to Hebert and Woodford (2018), Morris and Strack (2019), and Zhong

(2019), who adopt the posterior-based approach to study optimal stopping problems under RI with

general information cost functions in the continuous-time setup.3 Unlike their papers, ours is the

first to study optimal control problems under RI where the concavity of the objective function is

important for the optimality of the first-order conditions.

Most existing work on RI has focused on models with a continuous choice set, which are typi-

cally set up in the linear-quadratic-Gaussian framework (e.g., Peng and Xiong (2006), Luo (2008),

Mackowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao

(2019), and Miao, Wu, and Young (2019)). Woodford (2009) is the first paper that studies a dy-

namic binary choice problem under RI (the problem of a firm that decides each period whether to

reconsider its price). Jung et al (2018) show that rationally inattentive agents can constrain them-

selves voluntarily to a discrete choice set even when the initial choice set is continuous. See Sims

(2011) and Mackowiak, Matejka and Wiederholt (2018) for surveys and references cited therein.

2See Mattsson and Weibull (2002) and Fudenberg and Strzalecki (2015) for related models.3The posterior-based approach is often applied in the Bayesian persuasion literature. See Kamenica (2018) for a

survey and the references cited therein.

4

Page 5: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

2 Model

2.1 Setup

Consider a T -period decision problem with T ≤ ∞ and time is denoted by t = 1, 2, ..., T. Uncertainty

is represented by a discrete finite state space X ≡ {1, 2, ...,M} and a prior distribution µ1 ∈ ∆ (X) ,

where we use ∆ (Z) to denote the set of (probability) distributions on any finite set Z and ∆ (Y |Z)

to denote the set of conditional distributions on any finite set Y given any z ∈ Z.4 We also use a

bold case letter to denote any random variable such as x with its realization denoted by a normal

case letter x.

The decision maker (DM) makes choices from a finite action set denoted by A satisfying |A| ≥ 2.

We can allow the action set A to depend on the current state as in the literature on Markov decision

processes (Rust (1994) and Puterman (2005)), without affecting our key results but complicating

notation. The state transition kernel is given by π (xt+1|xt, at) , which defines the probability of

any state xt+1 ∈ X given any state xt ∈ X and any action at ∈ A for t ≥ 1. SSM (2017) show that

one can redefine the state space so that the state transition kernel is independent of the action.

We allow such dependence explicitly so that our model is more flexible in applications and is also

consistent with the literature on Markov decision processes (Rust (1994) and Puterman (2005)).

The DM receives flow utilities that depend on the current states and actions only. The period

utility function is given by a bounded function u : X × A → R. For the finite-horizon case with

T < ∞, we allow u to be time dependent and include a terminal utility function uT+1 : X → R.

SSM (2017) allow u to depend on the entire history of states and actions, which can generate

history-dependent solutions.

Prior to choosing an action in any period t, the DM can acquire costly information about the

history of the state xt, where we use xt to denote the history {x1, x2, ..., xt} and xtk to denote the

history {xk, xk+1, ..., xt} for k < t. More accurate information will lead to better choices, but are

more costly, with information costs to be discussed later. As MM (2015) and SSM (2017) show,

we do not need to model the endogenous choice of the information structure separately. Instead

we can reformulate the problem in which the DM makes stochastic choices and signals correspond

to actions directly. As CD (2013, 2015) argue, we can also identify signals with the corresponding

posteriors. Thus we will focus on the model with stochastic choices. See Lemma 3 in Appendix D

for a formal discussion.

Define a (state-dependent) choice rule as a sequence of conditional probability distributions{pt(at|xt, at−1

)∈ ∆

(A|Xt ×At−1

): all

(xt, at−1

), 1 ≤ t ≤ T

}.

The joint distribution of the state and action trajectories is denoted by{µt+1

(xt+1, at

)}, which

4As convention we define a conditional probability P (C|B) = P (C ∩B) /P (B) whenever P (B) > 0; otherwise,set P (C|B) = 0, which does not affect our analysis, but simplifies notation.

5

Page 6: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

is uniquely determined by the initial state distribution µ1 ∈ ∆ (X) , the state transition kernel

π (xt+1|xt, at) , and the preceding choice rule by a recursive formula

µt+1

(xt+1, at

)= π (xt+1|xt, at) pt

(at|xt, at−1

)µt(xt, at−1

)(1)

for any t ≥ 1. Set a0 = ∅ so that p1

(a1|x1, a0

)= p1 (a1|x1) and µ1

(x1, a0

)= µ1 (x1) .

Given a joint distribution µt+1

(xt+1, at

), we can compute the predictive distribution µt

(xt|at−1

),

the posterior distribution µt(xt|at

), and the distribution of an action conditional on a history of

actions qt(at|at−1

). Set µ1

(x1|a0

)= µ1 (x1) and q1

(a1|a0

)= q1 (a1) . Following SSM (2017), we

call qt(at|at−1

)a conditional default (choice) rule and the marginal distribution qt

(at)

of at an

unconditional default rule. The predictive distributions and posteriors satisfy

µt(xt|at−1

)=

∑at

qt(at|at−1

)µt(xt|at

), t ≥ 1, (2)

µt+1

(xt+1|at

)=

∑xt

π (xt+1|xt, at)µt(xt|at

), t ≥ 1. (3)

Equation (2) shows that the predictive distribution µt(xt|at−1

)can be computed by marginalizing

the posterior µt(xt|at

)over qt

(at|at−1

). Equation (3) shows that the future predictive distribution

µt+1

(xt+1|at

)can be transitioned from the current posterior µt

(xt|at

)using the transition kernel

π.

Conversely, starting from µ1 (x1) , given sequences of{µt(xt|at

)}and

{qt(at|at−1

)}, we can

compute µt(xt|at−1

)using (2) and use Bayes rule to derive the choice rule

pt(at|xt, at−1

)=µt(xt|at

)qt(at|at−1

)µt (xt|at−1)

, t ≥ 1. (4)

Then we can determine the joint distribution µt+1

(xt+1, at

)recursively by (1).

The joint distribution µT+1 ∈ ∆(XT+1 ×AT

)constructed above induces an expected dis-

counted utility value

E

[T∑t=1

βt−1u (xt,at) + βTuT+1 (xT+1)

],

where β ∈ (0, 1) denotes the discount factor.

2.2 Information Cost

In a static setup we follow CD (2013) and CDL (2019a,b) to define a UPS information cost function

as follows

CH (µ, µ (·|·) , q) ≡ H (µ)−∑a

q (a)H (µ (·|a)) ,

where H : ∆ (X) → R+ is a concave function (called generalized entropy), µ ∈ ∆ (X) is a prior

distribution, µ (·|·) ∈ ∆ (X|A) is a posterior, and q ∈ ∆ (A) is a marginal distribution that satisfies

6

Page 7: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

µ (x) =∑

a q (a)µ (x|a). The term H (µ) measures the amount of prior uncertainty and the term∑a q (a)H (µ (·|a)) measures the amount of uncertainty after acquiring information a. The concav-

ity of H implies CH (µ, µ (·|·) , q) ≥ 0 and the value of CH(µ, µ(·|·), q) represents the magnitude of

uncertainty reduction by observing information a about the state x. The following specifications

of H are interesting:

• Shannon entropy: H (ν) = −∑

x ν (x) ln ν (x), with limν(x)↓0 ν (x) ln ν (x) = 0.

• Weighted entropy (Belis and Guiasu (1968)):

H (ν) = −∑x

w (x) ν (x) ln ν (x) ,

where the weighting function satisfies w (x) ≥ 0 and∑

xw (x) = 1.

• Tsallis entropy (Havrda and Charvat (1967) and Tsallis (1988)):

H (ν) =1

σ − 1

∑x

ν (x)(

1− ν (x)σ−1), 0 < σ 6= 1.

• Renyi entropy (Renyi (1961)):

H (ν) =1

1− αln

(∑x

ν (x)α), α ∈ (0, 1) .

• Cressie-Read entropy (Cressie and Read (1984)):

H(ν) =∑x

ην(x)− ν(x)η

η(η − 1), 0 < η 6= 1. (5)

Shannon entropy is obtained as the limit (up to a constant) when σ → 1, α → 1, and η → 1.

The Tsallis entropy and Cressie-Read entropy cost functions are observationally equivalent up to

a scaling factor. Formally, the Tsallis entropy cost function with parameter σ is the same as the

Cressie-Read cost function with parameter η = σ multiplied by σ.

Figure 1 plots the Cressie-Read cost functions CH (µ, µ (·|·) , q) against µ (1|1) and their first

derivatives in the symmetric two-state two-action case with µ (1) = 0.5 and q (1) = 0.5 for different

values of η. Cost functions are convex. Functions with η > 1 have finite first derivatives as the

probability of state 1 tends to zero or one. This means that the marginal cost of information

does not go to infinity, so that the DM may choose to be fully informed. The marginal cost of

information increases as η decreases for µ (1|1) > 0.5.

In our dynamic setup, for the predictive distribution (prior belief) µt(·|at−1

)given history at−1,

we define the conditional information cost in period t of acquiring information at about the state

xt as

CH(µt(·|at−1

), µt(·|·, at−1

), qt(·|at−1

))= H

(µt(·|at−1

))−∑at

qt(at|at−1

)H(µt(·|at)),

7

Page 8: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

0 0.5 1

(1|1)

0

0.5

1

1.5

2

2.5

3Cressie-Read Cost Function

=0.3=0.5=1=1.3=2

0 0.5 1

(1|1)

-40

-30

-20

-10

0

10

20

30

40Marginal Cost

=0.3=0.5=1=1.3=2

Figure 1: Cressie-Read cost functions and their derivatives for different values of η for the symmetrictwo-state two-action case.

where µt(·|at−1

), µt

(·|·, at−1

), and qt

(·|at−1

)satisfy (2). The unconditional information cost in

period t of acquiring information at about the state xt is defined as

I(xt; at|at−1

)≡∑at−1

qt−1

(at−1

)CH

(µt(·|at−1

), µt(·|·, at−1

), qt(·|at−1

)), (6)

where

qt−1

(at−1

)= q0

(a0)q1

(a1|a0

)q2

(a2|a1

)· · · · · qt−1

(at−1|at−2

), q0

(a0)≡ 1.

The discounted information cost of acquiring information aT about xT is given by

T∑t=1

βt−1I(xt; at|at−1

),

where the sequence of prediction distributions{µt(·|at−1

)}satisfies (3). The Shannon mutual

information between aT and xT is the special case where β = 1 and H is the Shannon entropy

functon.

2.3 Decision Problem

We now formulate the dynamic discrete choice problem under RI as follows:

Problem 1 (posterior-based dynamic RI problem)

max E

[T∑t=1

βt−1u (xt,at) + βTuT+1 (xT+1)

]− λ

T∑t=1

βt−1I(xt; at|at−1

), (7)

8

Page 9: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

where the choice variables are sequences of distributions{µt(xt|at

)}and

{qt(at|at−1

)}for t ≥ 1

satisfying (2) and (3). Here I(xt; at|at−1

)is given by (6), and the expectation is taken with respect

to the joint distribution induced by π,{µt(xt|at

)}, and

{qt(at|at−1

)}.

Unlike the choice-based approach of MM (2015) and SSM (2017), the posterior-based approach

adopts posteriors instead of choice probabilities as a choice variable. The parameter λ > 0 measures

the shadow price of information in utility units. When λ = 0, the problem is reduced to the standard

Markov decision process formulation described in Puterman (2005) and Rust (1994). When λ > 0,

there is a tradeoff between information acquisition and utility maximization. Acquiring more precise

information about the state of the system helps the DM make a better choice. But this causes the

control actions to be statistically more dependent on the state, which generates a larger information

cost.

3 Preliminaries and Basic Intuition

In this section we first present the solution in the static case related to CD (2013), MM (2015), and

CDL (2019a). We then show that the choice-based approach of MM (2015) and SSM (2017) do not

work for the general UPS cost functions. Finally we study the two-period case and illustrate the

difficulty of the dynamic model and our solution approach.

3.1 Static Case

When T = 1 and uT+1 = 0, we obtain the following static problem according to the posterior-based

approach:

Problem 2 (static RI problem with UPS cost)

V (µ) ≡ maxq∈∆(A),µ(·|·)∈∆(X|A)

E [u (x,a)]− λCH (µ, µ (·|·) , q)

subject to

µ (x) =∑a

µ (x|a) q (a) , x ∈ X. (8)

Following CD (2013) and CDL (2019a), we rewrite this problem as

V (µ) ≡ maxq∈∆(A),µ(·|·)∈∆(X|A)

∑a

q (a)NaH (µ (·|a)) , V (µ) = V (µ)− λH (µ) (9)

subject to (8), where NaH (µ (·|a)) denotes the net utility of action a defined as

NaH (µ (·|a)) ≡

∑x

µ (x|a)u (x, a) + λH (µ (·|a)) . (10)

9

Page 10: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Notice that NaH (µ (·|a)) is concave in µ (·|a), but the problem in (9) is not jointly concave in q

and µ (·|·) due to the cross product term as pointed out by CD (2013). Thus one cannot simply

use the Kuhn-Tucker conditions to solve this problem. CD (2013) instead propose a geometric

approach from the convex analysis and derive necessary and sufficient conditions for optimality.

We first state some properties of the solution. Proofs of all results in the main text are collected

in Appendix A.

Proposition 1 Consider Problem 2. (i) The optimal posteriors µ (·|a) for all chosen actions a

with q (a) ∈ (0, 1) are independent of the prior µ ∈ ∆ (X) in the convex hull of these posteriors.

(ii) The optimal payoff for the static RI problem is given by

V (µ) = V (µ)− λH (µ) =∑x

µ (x) V (x)− λH (µ) , (11)

where V (x) is independent of the prior µ ∈ ∆ (X) in the convex hull of the optimal posteriors µ (·|a)

for all chosen actions a with q (a) ∈ (0, 1) . (iii) V (µ) is concave in µ and for x = 1, . . . ,M − 1

and µ(x) ∈ (0, 1),∂V (µ)

∂µ(x)= V (x)− V (M). (12)

For the Shannon entropy case, V (µ) is convex in µ.

Part (i) is the LIP property discovered by CD (2013). Part (ii) can be best understood using

the geometric approach of CD (2013). Specifically, the optimal posterior µ (·|a) is the tangent point

of the net utility associated with the chosen action a and V (x) satisfies

V (µ) =∑a

q (a)NaH (µ (·|a)) =

∑x

V (x)µ (x) ,

at the optimum. The value V (µ) is the height above µ (x) of the convex hull connecting NaH (µ (·|a))

for all chosen actions a. The optimal posterior µ (·|a) is the tangent point of V (µ) and NaH(µ(·|a))

for each a with q(a) ∈ (0, 1). The value V (x) is the height of the hyperplane containing this convex

hull at the point with µ (x) = 1 and µ (x′) = 0 for all x′ 6= x. This value is independent of the prior

µ in that convex hull. This result does not appear in the literature and is critical for the analysis

of the dynamic model. Notice that we need at least two chosen actions to form a convex hull. If

there is only one chosen action a, then q (a) = 1 and the posterior is the same as the prior. In this

case the convex hull is a degenerate singleton.

Figure 2 is similar to Figure 5 of CDL (2019a) in the case with two states {x, x′} and two actions

{a, b} . Net utilities are represented by the two solid curves. The concavification V (µ) is the concave

envelope of these two curves. The optimal posteriors µ (·|a) and µ (·|b) are given by the tangent

points at which the hyperplane supports the two net utility functions. The value V (x) is given by

10

Page 11: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Utility

V (x)

V (x′)

u(b, x′)

u(a, x)

0 1x′Prob. state

−H(µ)

u(b, x)

u(a, x′)

µ(x′|a) µ(x′) µ(x′|b)

Na(µ(·|a))

N b(µ(·|b))V (µ)

Figure 2: The net utility function and concavification.

the height of the hyperplane at the point with µ (x) = 1. Both the optimal posteriors, V (x) , and

V (x′) are invariant to changes of µ (x′) within the interval (µ (x′|a) , µ (x′|b)) . If µ (x′) ∈ (0, µ (x′|a)],

then q (a) = 1 and µ (x′|a) = µ (x′) . If µ (x′) ∈ [µ (x′|b) , 1], then q (b) = 1 and µ (x′|b) = µ (x′) .

Part (iii) of Proposition 1 shows that V (µ) is a concave function because it is the concave

envelope of net utilities. It is also differentiable and satisfies an envelope condition (see Corollary 2

of CD (2013)). It is not clear whether V (µ) is concave as it is equal to the difference of two concave

functions by (11). In fact we can show that it is convex if H is the Shannon entropy function. This

issue poses a difficulty when solving the dynamic RI problem.

To derive a tractable characterization and facilitate numerical solutions, we focus on the fol-

lowing additive class of generalized entropy:

Assumption 1 Let H satisfy

H(ν) =∑x

h(ν (x)), ν ∈ ∆ (X) , (13)

where h is a differentiable concave function defined on [0, 1].

Shannon entropy, weighted entropy, Tsallis entropy, and Cressie-Read entropy all satisfy As-

sumption 1, but Renyi entropy violates it. Under this assumption, we apply Lemma 3 of CD (2013)

to derive the following necessary and sufficient conditions:

11

Page 12: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Proposition 2 Suppose that Assumption 1 holds. Then the pair of µ (·|·) and q is optimal for

Problem 2 if and only if: (i) Equation (8) holds. (ii) There exists a function V : X → R such that

for any chosen action a ∈ A with q (a) > 0 and for any x ∈ X,

V (x) = u(x, a) + λh′(µ(x|a)) + λf(µ(·|a)), (14)

where

f(ν) ≡∑x

[h(ν (x))− ν (x)h′(ν (x))

], ν ∈ ∆ (X) . (15)

(iii) For any unchosen action b ∈ A with q (b) = 0 and µb ∈ ∆ (X) such that

u (x, b) + λh′(µb (x)

)−[u (M, b) + λh′

(µb (M)

)]= V (x)− V (M) , (16)

for x = 1, 2, ...,M − 1, we have

∑x

I

(V (x)

λ− u(x, b)

λ− f(µb)

)≤ 1. (17)

Here the function I : h′ ([0, 1])→ [0, 1] is defined as I(y) = h′−1(y) where h′ ([0, 1]) is the image of

[0, 1] under h′. Moreover, the function V (x) defined in (14) satisfies (11).

Condition (ii) shows that the right-hand side of equation (14) is independent of any chosen action

a and hence can be defined as a function V (x). Condition (iii) is a critical sufficient condition for

optimality and helps determine the consideration set (CDL (2019a)).

To understand this proposition, consider the special case of Shannon entropy. Then we have

h (ν (x)) = −ν (x) ln ν (x) and f (ν) = 1 for any ν ∈ ∆ (X) . Equation (14) becomes

V (x) = u(x, a)− λ ln(µ(x|a)) if q (a) > 0,

which gives the ILR property of CD (2013). Solving yields

µ(x|a) = exp

[− V (x)

λ+u(x, a)

λ

], if q (a) > 0.

Plugging this equation into (8) yields

V (x) = −λ ln

[µ(x)∑

a q(a) exp (u (x, a) /λ)

].

We then obtain

µ(x|a) =µ(x) exp (u (x, a) /λ)∑a′ q (a′) exp (u (x, a′) /λ)

if q (a) > 0.

By (11) we can derive the value function

V (µ) = λ∑x

µ (x) ln

[∑a

q(a) exp

(u(x, a)

λ

)].

12

Page 13: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

It can be easily shown that condition (iii) is equivalent to the following condition in CD (2013)

and CDL (2019a) in the Shannon entropy case:∑x

µ (x|a) exp (u (x, b) /λ)

exp (u (x, a) /λ)=∑x

[µ(x) exp (u (x, b) /λ)∑a′ q (a′) exp (u (x, a′) /λ)

]≤ 1,

for any chosen a with q (a) > 0 and any unchosen b ∈ A.Motivated by the Arimoto-Blahut algorithm, we use the following algorithm to solve for the

optimal q (·) and µ (·|·) :

1. Initialize f(a) ∈ R and q ∈ ∆ (A) with q(a) > 0 for all a ∈ A.

2. For any x, use the Newton method to solve for V (x) that satisfies the equation:5

µ(x) =∑a

q(a)I

(V (x)

λ− u(x, a)

λ− f(a)

). (18)

3. For any a ∈ A and x ∈ X, compute

µa(x) = I

(V (x)

λ− u(x, a)

λ− f(a)

). (19)

4. Update f(a) via

f+(a) =∑x

[h(µa(x))− µa(x)h′(µa(x))

]→ f (a) .

5. Update q(a) via

q+(a) =∑x

µa(x) q(a)→ q (a) . (20)

6. Go back to step 2, until(q+, f+

)converges to

(q, f).

7. Find µb ∈ ∆ (X) that satisfies (16). Check whether (17) is satisfied, where V (x) is the

converged value obtained in Step 6.

When a fixed point q+ = q is obtained, either q+(a) = q(a) ∈ (0, 1] or q+(a) = q(a) = 0. In the

first case, action a is chosen and hence∑

x µa(x) = 1 by (20). That is, µa is the optimal posterior

distribution µ (·|a) . In the second case, action a is not chosen. Then∑

x µa(x) ≤ 1 for the iteration

in (20) to converge. Step 7 checks the sufficient condition (17) in Proposition 2. To implement this

step, we use (19) to substitute V in (16) and derive

h′(µb (x)

)− h′

(µb (M)

)= h′

(µb (x)

)− h′

(µb (M)

), x = 1, ...,M − 1.

We use these M − 1 equations together with∑M

x=1 µb (x) = 1 to numerically solve for µb ∈ ∆ (X) .

For the Shannon entropy case, we can derive a closed-form solution: µb(x) = µb(x)/[∑

x′ µb(x′)

].

5In applications we allow I to be well defined in extended domain R and extended image R+.

13

Page 14: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

3.2 Failure of the Choice-based Approach

MM (2015) and SSM (2017) solve RI problems with the Shannon entropy cost using the choice-

based approach. To understand this approach, we notice that Problem 2 for the Shannon entropy

case can be rewritten as

V (µ) = maxq∈∆(A),p∈∆(X|A)

∑x,a

p (a|x)µ (x)

[u (x, a)− λ ln

p (a|x)

q (a)

], (21)

subject to

q (a) =∑x

p (a|x)µ (x) , a ∈ A. (22)

Let F (p, q) denote the objective function in (21). We can verify that F (p, q) is jointly concave in

(p, q) . Blahut (1972, Theorem 4) establishes the following result:

Lemma 1 Let p ∈ ∆ (A|X) be fixed. Then maxq∈∆(A) F (p, q) is a concave optimization problem

and the optimal solution is given by q (a) =∑

x µ (x) p (a|x) .

This lemma implies that the static RI problem (21) is equivalent to the following unconstrained

optimization problem:

maxp∈∆(A|X), q∈∆(A)

F (p, q) . (23)

Taking first-order conditions with respect to p and q yields the choice-based characterization as in

MM (2015) and CDL (2019a). CDL (2019a) also provide sufficient conditions for optimality.

To illustrate why the choice-based approach may not work for general UPS cost functions, we

let H be the weighted entropy. Then the cost function becomes

CH(µ, µ(·|·), q) =∑x,a

w(x)q (a)µ (x|a) lnµ (x|a)

µ (x)=∑x,a

w(x)µ(x)p(a|x) lnp(a|x)

q(a).

Following the choice-based approach described above, we define

F (p, q) =∑a,x

µ(x)p(a|x)

[u(x, a)− λw(x) ln

p(a|x)

q(a)

].

One can check that Lemma 1 does not hold in general so that the static RI problem is not equivalent

to the unconstrained problem in (23) for general UPS cost functions. Similarly, Lemma 2 in SSM

(2017) also fails for general UPS cost functions in dynamic RI models. Thus the coordinate-wise

first-order conditions for p and q cannot be used to characterize the solutions to dynamic RI

problems.

14

Page 15: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

3.3 Two-Period Case

As a prelude for our dynamic analysis we study the two-period case with T = 2 and uT+1 = 0. We

write the objective function as

J (q1, µ1 (·|·) , q2, µ2 (·|·))

≡ E [u (x1,a1) + βu (x2,a2)]− λCH (µ1, µ1 (·|·) , q1)

−βλ∑a1

q1 (a1)CH (µ2 (·|a1) , µ2 (·|·, a1) , q2 (·|a1))

=∑a1,x1

q1 (a1) [µ1 (x1|a1)u (x1, a1) + λH (µ1 (·|a1))]− λH (µ1)

+β∑

a1,a2,x2

q1 (a1){q2 (a2|a1)

[µ2

(x2|a2

)u (x2, a2) + λH

(µ2

(·|a2))]− λH (µ2 (·|a1))

},

where

µ1 (x1) =∑a1

q1 (a1)µ1 (x1|a1) , x1 ∈ X, (24)

µ2 (x2|a1) =∑a2

µ2

(x2|a2

)q2 (a2|a1) , (25)

µ2 (x2|a1) =∑x1

π (x2|x1, a1)µ1 (x1|a1) , a1 ∈ A, x2 ∈ X. (26)

Problem 3 Choose q1 ∈ ∆ (A) , q2 (·|·) ∈ ∆ (A|A) , µ1 (·|·) ∈ ∆ (X|A) , and µ2 (·|·) ∈ ∆(X|A2

)to

solve

max J (q1, µ1 (·|·) , q2, µ2 (·|·))

subject to (24), (25), and (26).

We solve Problem 3 by dynamic programming using the prior/predictive distribution as the

state variable. First consider the RI problem in period 2 conditional on the history of chosen action

a1 with q1 (a1) > 0 :

V2 (µ2 (·|a1)) = maxq2(·|a1),µ2(·|·,a1)

∑a2

q2 (a2|a1)Na2H

(µ2

(·|a2))− λH (µ2 (·|a1)) (27)

subject to (25), where we define the net utility as

Na2H

(µ2

(·|a2))

=∑x2

µ2

(x2|a2

)u (x2, a2) + λH

(µ2

(·|a2)).

The choice variables are µ2 (·|·, a1) ∈ ∆ (X|A) and q2 (·|a1) ∈ ∆ (A) . Taking the predictive distri-

bution µ2 (·|a1) as the prior at history a1, we view this problem as a static RI problem and apply

Lemma 3 in CD (2013) to characterize the solution in period 2.

15

Page 16: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

As in Proposition 1, we can show that

V2 (µ2 (·|a1)) = V 2 (µ2 (·|a1))− λH (µ2 (·|a1)) =∑x2

µ2 (x2|a1) V2 (x2|a1)− λH (µ2 (·|a1)) , (28)

where V 2 (µ2 (·|a1)) is concave in µ2 (·|a1).

By dynamic programming, the problem in period 1 is to choose µ1 (·|·) ∈ ∆ (X|A) and q1 (·) ∈∆ (A) to solve:

V1 (µ1) = max∑a1,x1

q1 (a1) [µ1 (x1|a1)u (x1, a1) + λH (µ1 (·|a1))]− λH (µ1) (29)

+β∑a1

q1 (a1)V2 (µ2 (·|a1))

subject to (24) and (26). The link between the problems in the two periods is through the predictive

distribution {µ2 (x2|a1)} that satisfies (26).

Substituting (28) into (29) yields

V1 (µ1) = maxq1,µ1(·|·)

∑a1

q1 (a1)Na1G (µ1 (·|a1))− λH (µ1) , (30)

where we define the net utility associated with action a1 as

Na1G (µ1 (·|a1)) ≡

∑x1

µ1 (x1|a1)u (x1, a1) + βV 2 (µ2 (·|a1)) + λGa1 (µ1 (·|a1)) .

In the above equation, we define

Ga (ν) ≡ H (ν)− βH

(∑x

π (·|x, a) ν (x)

), a ∈ A, ν ∈ ∆ (X) . (31)

If π (·|x, a) is independent of a, then Ga (ν) will be independent of a.

By Proposition 1, V 2 (µ2 (·|a1)) is concave in µ2 (·|a1), and thus it is also concave in µ1 (·|a1) by

(26). Thus if Ga1 (µ1 (·|a1)) is concave in µ1 (·|a1) , then so is Na1G . For the posterior-based approach

to work, we impose the following condition:

Assumption 2 For any β ∈ (0, 1) and a ∈ A, the function Ga (ν) defined in (31) is concave in

ν ∈ ∆ (X) .

This assumption is satisfied for the Shannon entropy function (see the proof of Corollary 1 in

Appendix A). It may not be satisfied for all generalized entropy functions because the difference of

two concave functions may not be concave. Under Assumption 2, we can then view the problem

in (30) as a static RI problem and apply Lemma 3 in CD (2013) to characterize the solution. In

Appendix B, we will provide a tractable characterization using Propostion 2 under Assumptions 1

and 2. We will also use this result to study dynamic RI problems with any horizon by breaking

down such problems into a sequence of static problems.

16

Page 17: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

4 Main Results

In this section we analyze dynamic RI problems using dynamic programming, provide necessary and

sufficient conditions for optimality, and describe a numerical algorithm to solve these conditions.

4.1 Dynamic Programming

To study Problem 1, let V T (µ1) denote the maximized value in (7) for T < ∞. Choosing the

predictive/prior distribution as the state variable, we obtain a value function V T : ∆ (X)→ R. For

T =∞, we simply use V (·) to denote the corresponding value function.

Endow ∆ (X) with the weak topology. Let V denote the set of all continuous functions on

∆ (X) . Then V is a Banach space. Define a Bellman operator T on this space as

T v (µ) = maxµp∈∆(X|A),q∈∆(A)

∑x,a

q (a)µp (x|a)u (x, a)− λCH (µ, µp, q) + β∑a

q (a) v(µ′ (·|a)

)(32)

subject to

µ (·) =∑a

q (a)µp (·|a) , (33)

µ′ (·|a) =∑x

π (·|x, a)µp (x|a) . (34)

By the principle of optimality, V and {V s}Ts=1 satisfy the Bellman equations V = T V, V s = T V s−1,

and

V 0 (µ) =∑x

µ (x)uT+1 (x) for T <∞.

Proposition 3 For Problem 1 with T = ∞, there exists a unique function V on V that satisfies

the Bellman equation V = T V. Moreover, limT→∞ VT (µ) = V (µ) for any µ ∈ ∆ (X) .

Let the policy functions for the infinite-horizon Bellman equation be gq : ∆ (X) → ∆ (A) and

gp : ∆ (X)→ ∆ (X|A) .6 Then we can write the solution as q = gq (µ) and µp = gp (µ) . The solution

to the sequence problem in (7) can be generated as follows: For t = 1,

q1 (a1) = gq (µ1) (a1) , µ1 (x1|a1) = gp (µ1) (x1|a1) ,

µ2 (x2|a1) =∑x1

π (x2|x1, a1) gp (µ1) (x1|a1) .

For any t ≥ 2,

qt(at|at−1

)= gq

(µt(·|at−1

))(at) , µt

(xt|at

)= gp

(µt(·|at−1

))(xt|at) ,

6If there are multiple solutions to the Bellman equation, we will select any one from the policy correspondence.

17

Page 18: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

µt+1

(xt+1|at

)=∑xt

π (xt+1|xt, at) gp(µt(·|at−1

))(xt|at) .

Both the preceding infinite- and finite-horizon dynamic programming problems are difficult to

analyze. Like the special cases studied in Section 3, the optimization problems are not concave

and the continuation value may not be concave. Thus it is also difficult to solve the dynamic

programming problems numerically. In particular, the standard value function iteration method is

inefficient and inaccurate because one has to discretize the state space and optimize over a feasible

set of µp (·|a) and q (a) for all a that satisfy constraint (33).

The following result provides a characterization of the solution to the finite-horizon dynamic

RI problem in terms of first-order conditions.7

Proposition 4 Suppose that Assumptions 1 and 2 hold. Then the sequences of {qt(at|at−1)}Tt=1

and {µt(xt|at)}Tt=1 are the optimal solution to dynamic RI Problem 1 if and only if: (i) Equations

(2) and (3) hold. (ii) The following system of equations holds:

Vt(xt|at−1) = vt(xt, at) + λh′(µt(xt|at)) + λft(µt(·|at), at), (35)

vt(xt, at) = u(xt, at) + β

∑xt+1

π(xt+1|xt, at)[Vt+1(xt+1|at)− λh′(µt+1(xt+1|at))1{t<T}

], (36)

for any at with qt(at|at−1

)> 0 and for t = 1, . . . , T , where 1{t<T} is an indicator function taking

value 1 if t < T and 0 if t = T, and

ft(ν, a) ≡∑x

[h(ν (x))− ν (x) h′(ν (x))

]− β

∑x′

[h

(∑x

π(x′|x, a

)ν (x)

)−∑x

π(x′|x, a

)ν (x) h′

(∑x

π(x′|x, a

)ν (x)

)], (37)

for ν ∈ ∆ (X) , a ∈ A, and t = 1, ..., T − 1, with the terminal conditions VT+1(xT+1|aT ) =

uT+1 (xT+1) and fT (ν, a) ≡∑

x [h(ν (x))− ν (x) h′(ν (x))] . (iii) For any unchosen action at ∈ Awith qt

(at|at−1

)= 0 and µatt ∈ ∆ (X) such that[

vt(xt, at) + λh′(µatt (xt))

]−[vt(M,at) + λh′(µatt (M))

]= Vt

(xt|at−1

)− Vt

(M |at−1

),

for xt = 1, 2, ...,M − 1, we have

∑xt

I

(Vt(xt|at−1)

λ− vt(xt, a

t)

λ− ft(µatt , at)

)≤ 1, (38)

for t = 1, 2, ..., T. Moreover, the value function satisfies

V T−t+1(µt(·|at−1)) =∑xt

µt(xt|at−1)Vt(xt|at−1)− λH(µt(·|at−1)), t = 1, ..., T. (39)

7The infinite-horizon solution can be obtained by taking limits in Proposition 4 as T → ∞.

18

Page 19: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Like in the static case, equation (35) shows that Vt(xt|at−1) is independent of all chosen actions

at. Condition (iii) in this proposition is a necessary and sufficient condition for optimality, which

is critical for determining the consideration set in each period. This condition is similar to that of

Proposition 2 in the static case.

When h (µ (x)) = −µ (x) lnµ (x) , we can verify that Assumption 2 is satisfied so that we can

apply Proposition 4 to derive a dynamic logit characterization for the Shannon entropy case.

Corollary 1 The solution to dynamic RI Problem 1 with Shannon entropy satisfies the following

necessary and sufficient conditions: (i) For t = 1, . . . , T ,

µt(xt|at) =µt(xt|at−1) exp

(vt(xt, a

t)/λ)∑

btqt(bt|at−1) exp (vt(xt, bt, at−1)/λ)

, (40)

pt(at|xt, at−1

)=

qt(at|at−1) exp(vt(xt, a

t)/λ)∑

btqt(bt|at−1) exp (vt(xt, bt, at−1)/λ)

, (41)

qt(at|at−1

)=∑xt

pt(at|xt, at−1

)µt(xt|at−1

), (42)

vt(xt, at) = u(xt, at) + β

∑xt+1

π(xt+1|xt, at)Vt+1

(xt+1, a

t), (43)

Vt(xt, a

t−1)

= λ ln

(∑at

qt(at|at−1) exp(vt(xt, a

t)/λ))

, (44)

with the terminal condition VT+1

(xT+1, a

T)

= uT+1(xT+1). (ii) For any at ∈ A,

∑xt

µt(xt|at−1) exp(vt(xt, a

t)/λ)∑

btqt(bt|at−1) exp (vt(xt, bt, at−1)/λ)

≤ 1, (45)

with equality if qt(at|at−1) > 0. Moreover, the value function satisfies

V T−t+1(µt(·|at−1

))=∑xt

µt(xt|at−1

)Vt(xt, a

t−1), t = 1, 2, ..., T. (46)

SSM (2017) provide a characterization for the Shannon entropy case using the choice-based

approach. Here we adopt the posterior-based approach. As shown in Section 3.2, the choice-based

approach of SSM (2017) does not work for general UPS cost functions beyond Shannon entropy.

As discussed in Section 3.1, the first-order characterization is useful for numerical analysis.

Because the solution is generally history dependent for the dynamic case, numerical analysis is

complicated and even infeasible for the infinite-horizon case. To simplify dynamic solutions, we

turn to a Markovian characterization.

19

Page 20: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

4.2 Markovian Solution

We define the following notion of Markovian solutions.

Definition 1 An optimal solution to dynamic RI Problem 1 is Markovian if, for any two different

histories at−1 and{bt−2, at−1

}reached with positive probabilities and any t = 1, 2, ..., T , the implied

predictive distributions satisfy µt(xt|at−1) = µt(xt|at−1, bt−2).

Intuitively, the predictive distribution is the state variable in the posterior-based dynamic pro-

gramming problem. If this state variable is history independent, then the optimal solution must

also be history independent. We thus have the following result.

Proposition 5 For a Markovian solution to dynamic RI Problem 1, the posterior distribution

µt(xt|at

), the choice rule pt(at|xt, at−1), and the default rule qt(at|at−1) at any history at−1 reached

with positive probabilities take the form of µt(xt|att−1

), pt(at|xt, at−1), and qt(at|at−1), respectively,

for any t = 1, ..., T .

By Definition 1 and Proposition 5, we can simply write µt+1

(xt+1|at

), µt

(xt|at

), and qt

(at|at−1

)as µt+1 (xt+1|at), µt

(xt|att−1

), and qt (at|at−1) , respectively. We can then rewrite (2) and (3) as

µt(xt|at−1) =∑at

µt(xt|att−1)qt(at|at−1), (47)

µt+1 (xt+1|at) =∑xt

π (xt+1|xt, at)µt(xt|att−1

), (48)

for any at−1 leading to at with positive probabilities.

We say that a solution to the dynamic RI problem in (7) is interior if qt(at|at−1) > 0 for any

action at ∈ A and any history at−1, t ≥ 1. The following result shows that any interior solution is

Markovian.

Proposition 6 An interior solution to dynamic RI Problem 1 is Markovian, for which the optimal

posterior distribution µt(xt|at

)takes the form µt (xt|at) for any t ≥ 1.

Now we provide necessary and sufficient conditions for Markovian solutions, which may not be

interior, using the posterior-based approach. This result generalizes Proposition 3 and Lemma 6 of

SSM (2017) to allow corner solutions and UPS cost functions.

Proposition 7 Let Assumptions 1 and 2 hold. Then the sequences of{µt(xt|att−1

)}and {qt (at|at−1)}

are a Markovian solution to dynamic RI Problem 1 if and only if: (i) The following system of equa-

tions holds: (47), (48),

Vt(xt|at−1) = vt(xt, at) + λh′(µt(xt|att−1)) + λft(µt(·|att−1), at), (49)

20

Page 21: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

vt(xt, at) = u(xt, at) + β∑xt+1

π(xt+1|xt, at)[Vt+1(xt+1|at)− λh′(µt+1(xt+1|at))1{t<T}

], (50)

for any at with qt (at|at−1) > 0 and for t = 1, . . . , T , with the terminal condition VT+1(xT+1|aT ) =

uT+1 (xT+1), where 1{t<T} and ft are defined in Proposition 4. (ii) For any unchosen action at ∈ Awith qt (at|at−1) = 0 and µatt ∈ ∆ (X) such that[

vt(xt, at) + λh′(µatt (xt))]−[vt(M,at) + λh′(µatt (M))

]= Vt (xt|at−1)− Vt (M |at−1) ,

for xt = 1, 2, ...,M − 1, we have

∑xt

I

(Vt(xt|at−1)

λ− vt(xt, at)

λ− ft(µatt , at)

)≤ 1, (51)

for t = 1, 2, ..., T . Moreover, the value function satisfies

Vt(µt(·|at−1)) =∑xt

µt(xt|at−1)Vt(xt|at−1)− λH(µt(·|at−1)), t = 1, ..., T. (52)

When h (µ (x)) = −µ (x) lnµ (x), the above proposition reduces to the Shannon entropy case

studied by SSM (2017). We can easily modify Corollary 1 to state a Markovian logit solution. We

omit the details here. Notice that not every dynamic RI problem admits an optimal Markovian

solution. We have solved numerical examples to illustrate this point for the Shannon entropy case.

This result is available upon request.

4.3 Numerical Methods

The conditions presented in Proposition 7 are a system of nonlinear difference equations, which is

nontrivial to solve both analytically and numerically. To solve this system numerically, we extend

the forward-backward Arimoto-Blahut algorithm proposed by Tanaka, Sandberg, and Skoglund

(2018) to our dynamic RI model with UPS costs. We present the algorithm in Appendix C. Here

we sketch the key idea.

We classify the difference equations in Proposition 7 into two groups. Equation (48) forms the

first group, which characterizes the evolution of the state variable {µt+1 (xt+1|at)}, and equations

(47), (49), (50), and (37) form the second group, which characterizes{µt(xt|att−1

)}, {qt (at|at−1)} ,{

ft(µt(xt|att−1), at)}, {vt(xt, at)} , and {Vt(xt|at−1)}.

To facilitate numerical implementation, we replace (48) with the following equations:

µt+1(xt+1, at) =∑

xt,at−1

π(xt+1|xt, at)µt(xt|att−1)qt(at|at−1)qt−1(at−1), q0 (a0) = 1, (53)

µt+1(xt+1|at) =µt+1(xt+1, at)

qt(at), qt(at) =

∑xt+1

µt+1(xt+1, at) > 0. (54)

21

Page 22: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Notice that the two µt+1 (xt+1|at) defined in (54) and (48) may not be identical even if we use the

same notation.

Clearly, µt+1 (xt+1|at) defined in (48) satisfies (54). But conversely µt+1(xt+1|at) defined in (54)

may not satisfy (48) unless µt(·|att−1) is the same for any two different actions at−1 and bt−1 reaching

the same at. The latter property can be satisfied if the transition matrix π(·|·, a) is invertible for

any a ∈ A. Under this condition we can use (48) to easily prove that for a Markovian solution

µt(·|at, at−1) = µt(·|at, bt−1) for any two different actions at−1 and bt−1 reaching at.

If the solution for sequences of posteriors{µt(xt|att−1

)}and conditional default rules {qt (at|at−1)}

is known, equations (53) and (54) in the first group can be solved forward in time to obtain

a sequence of predictive distributions {µt+1 (xt+1|at)} . On the other hand, if the solution for

{µt+1 (xt+1|at)} is known, the equations in the second group, which can be viewed as Bellman

equations, can be solved backward in time. Each time we use the static algorithm described in

Section 3.1 to solve for{µt(xt|att−1

)}and {qt (at|at−1)} . We solve the two groups of equations it-

eratively until convergence and check whether (48) is satisfied. We choose (53) and (54) instead of

(48) at each iteration because (53) and (54) ensure µt+1 (xt+1|at) to be always history independent,

but (48) may generate µt+1 (xt+1|at) that depends on at−1 before convergence.

We will use the above algorithm to solve some numerical examples in the next section. Whenever

a Markovian solution exists, our algorithm will find such a solution. We can design a similar

algorithm for the history-dependent solution in Proposition 4. This algorithm becomes complicated

for long-horizon problems as the history increases with the horizon and becomes infeasible under

infinite horizon.

5 Applications

In this section we apply our results to a matching state problem often studied in the literature

(SSM (2017), CD (2013) and CDL (2019a)). This problem can be used to describe many economic

decisions, e.g., consumer choices, project selection, and job search. Suppose that X = A and the

utility function satisfies u (xt, at) = 1 if xt = at; and u (xt, at) = 0, otherwise. We assume that the

transition kernel is independent of actions in Section 5.1 as in SSM (2017) and allow it to depend on

actions in Section 5.2. For simplicity we also assume that |X| = |A| = 2 and µ1 (x1 = 1) = 0.5.8 We

adopt the Cressie-Read entropy in (5) and compare the solution with that in the Shannon entropy

case.9

8Our algorithm works for larger state and action spaces.9We verify that Assumptions 1 and 2 and all conditions in Proposition 7 are satisfied in all our numerical examples

in this section.

22

Page 23: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

5.1 Transition Kernel Independent of Actions

As in SSM (2017), we assume π (xt+1|xt, at) = γ whenever xt+1 6= xt for any at ∈ A. We use

this example to illustrate that rationally inattentive behavior exhibits status quo bias over a short

horizon, but not over an infinite horizon. Moreover, the infinite-horizon behavior exhibits inertia.

As a benchmark, the optimal solution for the case without information cost (λ = 0) is to choose

an action to match the state in each period.

With information cost λ > 0, we first consider the interior Markovian solution in the infinite-

horizon stationary case, in which pt (at|xt, at−1) , qt (at|at−1) , and µt (xt|at) do not depend on time.

By equations (2) and (3), we have

q (1|1)µ (2|1) + q (2|1)µ (2|2) = (1− γ)µ (2|1) + γµ (1|1) .

By payoff symmetry µ (1|1) = µ (2|2) and q (1|1) = q (2|2) . Then we obtain

q (at = 1|at−1 = 1) = q (at = 2|at−1 = 2) = 1− γ,

as long as µ (2|1) 6= µ (1|1) . By prior symmetry the initial default rule satisfies q1 (a1 = 1) = 1/2.

Using the forward-backward algorithm, we numerically solve for the whole transition path. We

can verify the above interior solution and find that there is no transition in this example. In

particular, the solution immediately reaches the stationary case in period 2.

Our solution above verifies part 1 of Proposition 5 in SSM (2017), which considers more general

payoff functions and transition probabilities in the Shannon entropy case. The DM’s choices exhibit

inertia. That is, when the exogenous state is more persistent, the DM’s choice behavior is also more

persistent. For our example, they have the same persistence 1− γ.

It is more interesting to analyze the finite-horizon solution. SSM (2017) study the two-period

case with Shannon entropy and their Proposition 4 shows that when γ is sufficiently small, q2 (1|1) =

q2 (2|2) = 1 and Pr (a1 = a2) = 1. That is, if the probability of changing states is sufficiently small,

the DM’s behavior exhibits status quo bias in the sense that he acquires information only in the first

period and relies on that information in both periods. Through extensive numerical experiments,

we find that this result does not hold in the infinite-horizon case. In particular, we always have the

interior solution described above for any γ ∈ (0, 1) given µ1 (x1 = 1) = 0.5.

The intuition behind the above result is the following. In the two-period case, when γ is

sufficiently small, the DM believes that any state in period 1 is more likely to remain the same in

period 2. Thus the DM does not want to acquire new information and just follows the first period

choice. However, when the horizon becomes longer, future states are more likely to switch. In

particular, the switching probability is given by 1− (1− γ)T , which increases to 1 for γ ∈ (0, 1) as

T → ∞. Thus it is more valuable to acquire new information when the decision horizon is longer.

23

Page 24: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

0 2 4 6 8 10

time

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Choice probability

=0.2, pt(1|1,1)

=0.2 pt(1|1,2)

=1 pt(1|1,1)

=1 pt(1|1,2)

=1.3 pt(1|1,1)

=1.3 pt(1|1,2)

0 2 4 6 8 10

time

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Default rule

=0.2 qt(1|1)

=0.2 qt(1|2)

=1 qt(1|1)

=1 qt(1|2)

=1.3 qt(1|1)

=1.3 qt(1|2)

Figure 3: Choice probabilities and default rules for different values of η. Parameter values areT = 10, µ1(0) = 0.5, π(xt+1|xt, at) = γ = 0.04 if xt+1 6= xt, β = 0.8, and λ = 1.

But when the decision horizon is sufficiently short, the DM will not acquire any information, e.g.,

in the terminal period.

Figure 3 illustrates the analysis above for different values of the curvature parameter η for the

Cressie-Read entropy. We set T = 10, µ1 (x1 = 1) = 0.5, λ = 1, γ = 0.04, u11 = 0, and β = 0.8. For

the baseline Shannon entropy case with η = 1, we find that q1 (a1 = 1) = 1/2 by symmetry and

qt (at = 2|at = 2) = qt (at = 1|at−1 = 1) =

0.96 for t = 2, 3, ..., 7,0.9858 for t = 8,1 for t = 9, 10.

Thus the status quo bias behavior occurs starting from period 9. The left panel of Figure 3

presents the paths of pt (at = 1|xt = 1, at−1 = 1) and pt (at = 1|xt = 1, at−1 = 2) . At time t = 1,

they are the same because a0 is not present. Then pt (at = 1|xt = 1, at−1 = 1) increases to 1

and pt (at = 1|xt = 1, at−1 = 2) decreases to zero, consistent with the inertia behavior shown in

Proposition 5 of SSM (2017). We also find that, when T → ∞, there is no terminal time and

qt (at = 1|at−1 = 1) = 0.96 = 1− γ for all t ≥ 2.

Next we consider solutions when η 6= 1. We find that the status quo bias behavior occurs earlier

when η is smaller. Moreover, the state dependent choice probabilities pt (at = xt|xt, at−1) decrease

as η decreases. The intuition follows from Figure 1 presented in Section 2.2. When η is smaller, the

marginal cost of information is larger so that the DM has less incentive to acquire new information.

Then the DM is more likely to make mistakes and stick to the old information.

Our analysis indicates that it is rational inattention combined with the short horizon that

generates status quo bias. This bias does not exist under infinite horizon. But the inertia behavior

exists in both finite- and infinite-horizon settings. Moreover, the timing of status quo bias depends

24

Page 25: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

on the marginal cost of information, which depends on the specification of the information cost

function.

5.2 Transition Kernel Depends on Actions

We now show that the results are very different when the state transition kernel depends on actions.

For simplicity, assume that π (xt+1|xt, at) = α ∈ [0, 1] if xt+1 = at, for t ≥ 1. That is, the probability

that the state in the next period confirms the current action is equal to α and is independent of

the current state. We use this example to show that status quo bias can persist in the long run

and confirmation bias and belief polarization can also arise.

Notice that the optimal solution in the case without information cost (λ = 0) is always to

choose an action to match the state in each period as in the previous subsection. Next consider

the two-period case with costly information acquisition (λ > 0).

Proposition 8 Consider the two-period RI model with Shannon entropy. Let u3 = 0, β = 1,

and π (xt+1|xt, at) = α ∈ [0, 1] whenever xt+1 = at. Let α∗ ≡ exp(1/λ)exp(1/λ)+1 and α∗∗ ≡ 1

exp(1/λ)+1 .

Then the solution satisfies q1 (1) = 1/2 and Pr (a2 = a1) = 1 for α > α∗ and q1 (1) = 1/2 and

Pr (a2 6= a1) = 1 for α < α∗∗. For α ∈ (α∗, α∗∗) , the solution is interior with

q2 (1|1) = q2 (2|2) =α (exp (1/λ) + 1)− 1

exp (1/λ)− 1.

This proposition shows that the status quo bias can emerge for reasons other than those dis-

cussed in the previous subsection. In particular, if α is sufficiently large, the DM believes that

there is a high probability that x2 confirms a1 and thus he does not reverse his decision. But if α

is sufficiently small, he reverses his decision.

Using numerical methods, we find that the solutions for any T > 2 are similar to those for

T = 2. This result is different from the case in which the state transition kernel is independent

of actions. In that case the status quo bias does not occur under infinite horizon because the

probability that the state will eventually switch is equal to 1. By contrast, for the model in this

subsection, the state transition probability is independent of the current state, but dependent on

the current action. If the probability that the state in the next period matches the current action

is sufficiently high, the DM will not reverse his initial decision in that Pr (at = a1) = 1 for all t > 1.

On the other hand, if this probability is sufficiently low, the DM will reverse his initial decision

forever in that Pr (at = a′1) = 1 for all t > 1 and a′1 6= a1.

We are unable to derive an analytical result similar to Proposition 8 for general UPS cost

functions. We thus solve numerical examples using Cressie-Read entropy. Figure 4 illustrates the

transition dynamics for different values of η. We set the parameter values T = 10, λ = 1, β = 0.8,

u11 = 0 and α = 0.7. We find that there is no transition and the solution becomes stationary from

25

Page 26: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

0 5 10

time

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Choice probability

=0.2, pt(1|1,1)

=0.2 pt(1|1,2)

=1 pt(1|1,1)

=1 pt(1|1,2)

=1.3 pt(1|1,1)

=1.3 pt(1|1,2)

0 5 10

time

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Default rule

=0.2 qt(1|1)

=0.2 qt(1|2)

=1 qt(1|1)

=1 qt(1|2)

=1.3 qt(1|1)

=1.3 qt(1|2)

Figure 4: Choice probabilities and default rules for different values of η. Parameter values areT = 10, µ1(0) = 0.5, π(xt+1|xt, at) = α = 0.7 if xt+1 = at, β = 0.8, and λ = 1.

the second period on. For this value of α, we have Pr (a2 = a1) = 1 for η = 0.2. For η = 1 and

1.3, the solutions are interior, but η = 0.2 generates a corner solution. The difference in η is also

reflected in the initial choice probabilities p1 (a1 = 1|x1 = 1) . Figure 4 shows that p1 (a1 = 1|x1 = 1)

declines as η decreases. For a smaller value of η, the marginal cost of information is larger and the

DM has less incentive to acquire new information. Thus the DM is more likely to make mistakes.

There is a positive feedback between beliefs and actions in the model of this subsection. When

the DM believes that the state in the next period is sufficiently likely to be consistent with the

DM’s current action, he will choose the same action in the next period in order to match the state.

In this case he acquires information only in period 1 and uses the same information in the future.

Even though the realized state in the future is different from his initial action, he still mistakenly

sticks to the initial chosen action because processing new information is costly.

The model here has implications for confirmation bias and belief polarization in the psychology

literature. Confirmation bias is the tendency to search for, interpret, favor, and recall information

in a way that confirms one’s preexisting beliefs or hypotheses. This behavior happens in our model

because the DM will stick to his initial choice if he entertains a strong belief that the future state

is likely to be consistent with his current action. If there are more individuals, belief polarization

may occur. Suppose that two individuals with the same prior about the states have different beliefs

about state transition probabilities. One believes the future state is more likely to be consistent

with the current action, and the other believes the opposite. Then after the same state is realized

over time, each one believes his own belief is correct. One individual will choose the same action

as the initial one. The other will always choose the action different from the initial one.

26

Page 27: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

6 Conclusion

We adopt the posterior-based approach to study dynamic RI problems and provide necessary and

sufficient conditions for optimal solutions for the additive class of UPS cost functions. We propose an

efficient algorithm to solve these conditions and apply our model to explain some behavioral biases.

Because the large class of cost functions considered in our paper can help explain some behavior

that violates the predictions of the RI models with the Shannon entropy cost, our approach will

find wide applications in dynamic settings.

27

Page 28: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Appendix

A Proofs

Proof of Proposition 1: (i) It follows from Corollary 1 of CD (2013).

(ii) Let V (x) be the height of the hyperplane at the point with µ (x) = 1 and µ (x′) = 0 for all

x′ 6= x. This hyperplane contains the convex hull spanned by the optimal posteriors µ (·|a) for all

chosen actions a with q (a) ∈ (0, 1) . Then V (x) is independent of the prior µ in that convex hull

by part (i). If there is only one chosen action a with q (a) = 1, then the convex hull is a singleton.

By construction, V (µ) =∑

x µ (x) V (x) . We obtain the desired result.

(iii) Let q∗i and µ∗i (·|·) be the optimal solution corresponds to any prior µi for i = 1, 2. Let

q (a) = θq∗1 (a) + (1− θ) q∗2 (a) for θ ∈ (0, 1) for any chosen action a. Then we can derive

θV (µ1) + (1− θ)V (µ2)

= θ∑a

q∗1 (a)NaH (µ∗1 (·|a)) + (1− θ)

∑a

q∗2 (a)NaH (µ∗1 (·|a))

=∑a

q (a)

[θq∗1 (a)

q (a)NaH (µ∗1 (·|a)) +

(1− θ) q∗2 (a)

q (a)NaH (µ∗2 (·|a))

]≤

∑a

q (a)NaH

(θq∗1 (a)

q (a)µ∗1 (·|a) +

(1− θ) q∗2 (a)

q (a)µ∗2 (·|a)

)≤ V (θµ1 + (1− θ)µ2) ,

where the first inequality follows from the concavity of the net utility NaH , and the second inequality

from the following: ∑a

q (a)

(θq∗1 (a)

q (a)µ∗1 (·|a) +

(1− θ) q∗2 (a)

q (a)µ∗2 (·|a)

)= θ

∑a

q∗1 (a)µ∗1 (·|a) + (1− θ) q∗2 (a)µ∗2 (·|a)

= θµ1 (·) + (1− θ)µ2 (·) .

Thus V (µ) is concave in µ.

If µ is in the convex hull of the optimal posteriors for at least two chosen actions, then V is

independent of µ in that convex hull. We then obtain (12). CDL (2019a) show that the set ∆ (X)

can be partitioned into sets of priors, each of which is associated with a given consideration set.

The derivative formula (12) applies to each set of priors and crosses boundaries of neighboring sets

continuously. We can also apply Corollary 2 of CD (2013) to derive (12).

Finally consider the Shannon entropy case. Then Problem 2 can be reformulated as in (21) and

(22) (see MM (2015)). Let θ ∈ (0, 1) , µ, µ′ ∈ ∆ (X) , and µ∗ = θµ + (1− θ)µ′. Let p∗ (a|x) and

28

Page 29: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

q∗ (a) be the associated optimal solution. Then

q∗ (a) =∑x

p∗ (a|x)µ∗ (x) = θq∗1 (a) + (1− θ) q∗2 (a) ,

where

q∗1 (a) =∑x

p∗ (a|x)µ (x) , q∗2 (a) =∑x

p∗ (a|x)µ′ (x) .

Since Shannon entropy is a concave function, we deduce that

V(θµ+ (1− θ)µ′

)=∑x,a

p∗ (a|x)µ∗ (x)

[u (x, a)− λ ln

p∗ (a|x)

q∗ (a)

]=

∑x,a

p∗ (a|x)µ∗ (x) [u (x, a)− λ ln p∗ (a|x)]− λH (q∗)

≤ θ∑x,a

p∗ (a|x)µ (x)

[u (x, a)− λ ln

p∗ (a|x)

q∗1 (a)

]+ (1− θ)

∑x,a

p∗ (a|x)µ′ (x)

[u (x, a)− λ ln

p∗ (a|x)

q∗2 (a)

]≤ θV (µ) + (1− θ)V

(µ′).

Thus V (µ) is convex. Q.E.D.

Proof of Proposition 2: Recall that

NaH (µ (·|a)) ≡

∑x

µ (x|a)u (x, a) + λH (µ (·|a)) . (A.1)

To apply Lemma 3 of CD (2013), we compute

∂NaH(µ(·|a))

∂µ(x|a)= u(x, a) + λh′(µ(x|a))− u(M,a)− λh′ (µ(M |a)) , (A.2)

for any x = 1, . . . ,M − 1, and

NaH(µ(·|a))−

M−1∑x=1

∂NaH(µ(·|a))

∂µ(x|a)µ(x|a) (A.3)

= λM∑x=1

[h(µ(x|a))− µ(x|a)h′(µ(x|a))

]+ u(M,a) + λh′(µ(M |a)).

We will show that conditions in Proposition 2 are equivalent to conditions (ED), (CT), and

(UB) in Lemma 3 of CD (2013). Condition (ED) is equivalent to

u(x, a)+λh′(µ(x|a))−u(M,a)−λh′(µ(M |a)) = u(x, b)+λh′(µ(x|b))−u(M, b)−λh′(µ(M |b)), (A.4)

for any chosen actions a, b, and any x = 1, . . . ,M − 1.

29

Page 30: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Condition (CT) is equivalent to

λ∑x

[h(µ(x|a))− µ(x|a)h′(µ(x|a))

]+ u(M,a) + λh′(µ(M |a)) (A.5)

= λ∑x

[h(µ(x|b))− µ(x|b)h′(µ(x|b))

]+ u(M, b) + λh′(µ(M |b)),

for any chosen actions a and b.

Condition (UB) is equivalent to

λ∑x

[h(µb(x))− µb(x)h′(µb(x))

]+ u(M, b) + λh′(µb(M))

≤λ∑x

[h(µ(x|a))− µ(x|a)h′(µ(x|a))

]+ u(M,a) + λh′(µ(M |a)),

(A.6)

for any chosen action a, any unchosen action b, and µb ∈ ∆(X) such that∂Nb

H(µb(·))∂µb(x)

=∂Na

H(µ(·|a))

∂µ(x|a)

for any x = 1, . . . ,M − 1 .

Define function f as in (15). Then we can show that conditions (A.4) and (A.5) are equivalent

to the following condition:

u(x, a) + λh′(µ(x|a)) + λf(µ (·|a)) = u(x, b) + λh′(µ(x|b)) + λf(µ (·|b)), (A.7)

for any x = 1, 2, ...,M and any chosen actions a and b.

For any chosen action a, the height of the hyperplane passing through NaH(µ(·|a)) at µ (·) is

given by

NaH(µ(·|a))−

M−1∑x=1

∂NaH(µ(·|a))

∂µ(x|a)(µ(x|a)− µ(x)).

By conditions (CT) and (ED) in Lemma 3 of CD (2013), the expression above is independent of

any chosen action a. The height of this hyperplane at the prior with µ (x) = 1 and µ (x′) = 0 for

x′ 6= x is

V (x) = NaH(µ(·|a))−

M−1∑y=1

∂NaH(µ(·|a))

∂µ(y|a)µ(y|a) +

∂NaH(µ(·|a))

∂µ(x|a), x = 1, . . . ,M − 1,

V (M) = NaH(µ(·|a))−

M−1∑y=1

∂NaH(µ(·|a))

∂µ(y|a)µ(y|a).

(A.8)

By (A.2), (A.3), and (A.7), we obtain (14). Conversely, let (14) hold. Because V (x) is independent

of chosen action, (A.7) holds so that conditions (CT) and (ED) in (A.4) and (A.5) hold. Thus

we have shown that condition (ii) in Proposition 2 is equivalent to conditions (CT) and (ED). By

construction, V (x) satisfies (11).

By (A.2) and the definition of f, condition (UB) in (A.6) is equivalent to

u(M, b) + λh′(µb(M)) + λf(µb)≤ u(M,a) + λh′(µ(M |a)) + λf(µ(·|a)), (A.9)

30

Page 31: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

where q (a) > 0 and µb ∈ ∆ (X) satisfies[u(x, b) + λh′(µb(x))

]−[u(M, b) + λh′(µb(M))

]=[u(x, a) + λh′(µ(x|a))

]−[u(M,a) + λh′(µ(M |a))

], (A.10)

for x = 1, ...,M − 1. Notice that (A.9) and (A.10) imply

u(x, b) + λh′(µb(x)) + λf(µb)≤ u(x, a) + λh′(µ(x|a)) + λf(µ(·|a)), (A.11)

for x = 1, 2, ...,M. Conversely, suppose that (A.10) holds but (A.9) fails for some chosen action a

and some action b ∈ A. Then we can check that (A.11) fails too. Thus we have shown that (A.11)

is equivalent to condition (UB) in (A.6) given (A.10).

By (A.11), (14), and the definition of I, we obtain

µb(x) ≥ I

(V (x)

λ− u(x, b)

λ− f(µb)

), x = 1, . . . ,M. (A.12)

Since∑

x µb(x) = 1, this inequality implies (17). Here µb satisfies (A.10), which is equivalent to

(16) using (14). Conversely, if (A.12) fails for some x, the previous argument using (A.10) shows

that it also fails for all other x. Hence (17) fails as well. Therefore (UB) is equivalent to condition

(iii) in Proposition 2. Q.E.D.

Proof of Proposition 3: We can easily verify the operator T satisfies the Blackwell sufficient

condition. Thus it is a contraction mapping. Since ∆ (X) is compact when endowed with the weak

topology, continuous functions on this space are bounded. Thus V is a Banach space. We can verify

that T maps a function in V into V by the theorem of the maximum. By the contraction mapping

theorem, there is a unique fixed point V ∈ V such V = T V. Moreover, lims→∞ T sV 0 = V for any

V 0 ∈ V. Thus limT→∞ VT = V. See Stokey, Lucas with Prescott (1989) for a reference of the cited

theorems here. Q.E.D.

Proof of Proposition 4: The sequence of value functions satisfies the dynamic programming

equations at any history at−1 reached with positive probabilities:

V T−t+1(µt(·|at−1

))= max

µt(·|·,at−1),qt(·|at−1)

∑xt,at

qt(at|at−1

)µt(xt|at

)u (xt, at)

−λCH(µt(·|at−1

), µt(·|·, at−1

), qt(·|at−1

))+ β

∑at

qt(at|at−1

)V T−t (µt+1

(·|at))

subject to (2) and (3) for t = 1, ..., T . In the last period we have a terminal condition

V 0(µT+1

(·|aT

))=∑xT+1

µT+1

(xT+1|aT

)uT+1 (xT+1) .

31

Page 32: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Starting from the last period T, we apply the analysis in Appendix B recursively by backward

induction. We can then derive Proposition 4. Here we only outline the key step and omit the

detailed derivation. Plugging equation

V T−t (µt+1

(·|at))

=∑xt+1

µt+1

(xt+1|at

)Vt+1

(xt+1|at

)− λH

(µt+1

(·|at))

into the above Bellman equation, we obtain

V T−t+1(µt(·|at−1

))= max

µt(·|·,at−1),qt(·|at−1)

∑xt,at

qt(at|at−1

)NatG

(µt(xt|at

))− λH

(µt(·|at−1

))where the net utility function is defined as

NatG

(µt(·|at))

=∑xt

µt(xt|at

)u (xt, at) + βV t+1

(µt+1

(·|at))

+ λGat(µt(·|at)).

By Assumption 2, Gat(µt(·|at))

is concave in µt(·|at). The concave envelope V t+1

(µt+1

(·|at))

=∑xt+1

µt+1

(xt+1|at

)Vt+1

(xt+1|at

)is concave in µt+1

(·|at)

by Proposition 1 and hence in the pos-

terior µt(·|at)

by (3). Thus the net utility function NatG

(µt(·|at))

is concave. We can then use

Proposition 2 to characterize the solution. Q.E.D.

Proof of Corollary 1: In the Shannon entropy case, h(µ(x)) = −µ(x) lnµ(x). We first verify

that Assumption 2 is satisfied so that we can apply Proposition 4. For any µ, µ ∈ ∆ (X) , define

Ga(µ, µ) as

Ga(µ, µ) =∑x1,x2

π(x2|x1, a)µ(x1) ln[µ(x2)]β

µ(x1).

Therefore,

Ga(ν) = Ga

(ν,∑x

π(·|x, a)ν(x)

).

Notice that Ga(µ, µ) is a convex combination of µ(x1) ln [µ(x2)]β

µ(x1) for all x1, x2 ∈ X. The expression

µ(x1) ln [µ(x2)]β

µ(x1) is a jointly concave function of µ(x1) and µ(x2) for any β ∈ (0, 1]. Therefore, Ga is

jointly concave in µ and µ.

For any θ ∈ [0, 1] and ν, ν ′ ∈ ∆ (X),

Ga(θν + (1− θ)ν ′

)= Ga

(θν + (1− θ)ν ′, θ

∑x

π(·|x, a)ν(x) + (1− θ)∑x

π(·|x, a)ν ′(x)

)

≥ θGa(ν,∑x

π(·|x, a)ν (x)

)+ (1− θ)Ga

(ν,∑x

π(·|x, a)ν ′ (x)

)= θGa(ν) + (1− θ)Ga(ν ′),

32

Page 33: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

where the inequality follows from the definition of a jointly concave function. Thus Assumption 2

is satisfied for Shannon entroy.

In the Shannon entropy case, we have ft(µt(·|at), at) = 1 − β for t = 1, . . . , T − 1, and

fT (µT(·|aT

), aT ) = 1, for any chosen aT . We obtain from (35) that

µt(xt|at) = exp

(− Vt(xt|a

t−1)

λ+vt(xt, a

t)

λ− β1{t<T}

),

where 1 is an indicator function. Using this equation and (2), we can solve for Vt(xt|at−1):

Vt(xt|at−1) = −λ ln

µt(xt|at−1)∑btqt(bt|at−1) exp

(vt(xt,bt,at−1)

λ − β1{t<T}

) .

Define vt(xt, at) = vt(xt, a

t)− λβ1{t<T} and define Vt(xt, a

t−1)

as in (44). Combining the previous

two equations, we confirm (40). Plugging the previous expression of Vt(xt|at−1) into (36), we

confirm the recursive relation (43) for vt. Equations (41) and (42) follow from the usual probability

rules. Inequality (45) follows from (38) and (46) follows from (39). Q.E.D.

Proof of Proposition 5: By Definition 1, at history at−1 reached with positive probabilities,

µt(·|at−1

)takes the form µt (·|at−1) . Then the DM solves the following Bellman equation:

Vt (µt (·|at−1))

= maxµt(·|·,at−1),qt(·|at−1)

∑xt,at

qt(at|at−1

)µt(xt|at

)u (xt, at) (A.13)

−λCH(µt (·|at−1) , µt

(·|·, at−1

), qt(·|at−1

))+ β

∑at

qt(at|at−1

)Vt+1

(µt+1

(·|at))

subject to

µt (xt|at−1) =∑at

qt(at|at−1

)µt(xt|at

), (A.14)

µt+1

(xt+1|at

)=

∑xt

π (xt+1|xt, at)µt(xt|at

), (A.15)

for t = 1, ..., T, with the terminal condition:

VT+1

(µT+1

(·|aT

))=∑x

µT+1

(x|aT

)uT+1 (x) .

As discussed in Section 4.1, the solution is a function of the prior/predictive distribution µt (xt|at−1)

independent of history at−2. Thus the optimal solution for qt(at|at−1

)and µt

(xt|at

)takes the form

qt (at|at−1) and µt(xt|att−1

).

33

Page 34: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

We can compute the state dependent choice probability:

pt+1(at+1|xt+1, at) =

µt+1(xt+1|at+1)qt+1(at+1|at)µt+1(xt+1|at)

=µt+1(xt+1|at+1, at)qt+1(at+1|at)

µt+1(xt+1|at)= pt+1(at+1|xt+1, at),

for any µt+1(xt+1|at) > 0. Q.E.D.

Proof of Proposition 6: The first-period predictive distribution is the prior µ1. The second-

period predictive distribution is µ2(·|a1). Because the solution is interior, q2(a2|a1) > 0 for any

a2 ∈ A. Then all predictive distributions µ2(·|a1) for different a1 are in the interior of the convex hull

spanned by optimal posteriors µ2(·|a2). By the LIP property of CD (2013), µ2(·|a2) is independent

of µ2(·|a1) and hence independent of a1. Thus µ2(·|a2) takes the form µ2(·|a2). The predictive

distribution in period t = 3 is determined by

µ3(x3|a2) =∑x2

π(x3|x2, a2)µ2(x2|a2),

which does not depend on a1. We can show that µt+1

(xt+1|at

)takes the form of µt+1 (xt+1|at)

using the same argument by induction. Thus an interior solution is Markovian. Moreover, the

optimal posterior µt(xt|at

)takes the form µt (xt|at) for any t ≥ 1. Q.E.D.

Proof of Proposition 7: For a Markovian solution, we directly apply Proposition 4 by replacing

any history-dependent variable with its history independent version to obtain conditions (i)-(iii) in

Proposition 7. Conversely, suppose that {qt(at|at−1)} and{µt(xt|att−1)

}satisfy these conditions.

Then the optimal posterior {µt(xt|at)} for problem (A.13) takes the form {µt(xt|att−1)} given the

prior belief µt(·|at−1) for any t ≥ 1. It then follows from (48) that the predictive distribution

µt+1(xt+1|at) is independent of the history at−1. Starting at t = 1, the prior is µ1 (·|a0) = µ1 (·) .By induction we can show that µt+1(xt+1|at) takes the form µt+1(xt+1|at) for any t ≥ 1. By

Definition 1 the solution is Markovian. Q.E.D.

Proof of Proposition 8: There are two types of solutions. By symmetry of the problem, we

first solve for a symmetric interior solution satisfying q1 (a1 = 1) = 1/2 and q2 (1|1) = q2 (2|2) = z.

Interior solutions are Markovian. By Corollary 1, we compute

V2 (1, 1) = V2 (2, 2) = λ ln [z exp (1/λ) + 1− z] ,

V2 (1, 2) = V2 (2, 1) = λ ln [(1− z) exp (1/λ) + z] ,

v1 (1, 1) = v1 (2, 2) = 1 + βαV2 (1, 1) + β (1− α) V2 (2, 1) ,

34

Page 35: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

v1 (1, 2) = v1 (2, 1) = βαV2 (2, 2) + β (1− α) V2 (1, 2) .

It follows from µ1 (1) = 1/2 that the DM’s initial value is given by

V1 =1

2λ ln

1

2[exp (v1 (1, 1) /λ) + exp (v1 (1, 2) /λ)]

+1

2λ ln

1

2[exp (v1 (2, 1) /λ) + exp (v1 (2, 2) /λ)]

= λ ln1

2[exp (v1 (1, 1) /λ) + exp (v1 (1, 2) /λ)] .

Thus maximizing V1 is equivalent to maximizing(ze

1λ + 1− z

)βα [(1− z) e

1λ + z

]β(1−α).

This is a concave function of z. The first-order condition gives

z =α (exp (1/λ) + 1)− 1

exp (1/λ)− 1.

Thus, if

α∗∗ ≡ 1

exp (1/λ) + 1< α <

exp (1/λ)

exp (1/λ) + 1≡ α∗,

then the optimal solution is interior z ∈ (0, 1) . If α ≥ α∗, the solution is at the corner z = 1. If

α ∈ [0, α∗∗] , the solution is at the other corner z = 0. We then obtain the desired result.

It remains to show that the corner solution in which q1 (1) = 1 is not optimal. Suppose that it

is optimal. Then we use q1 (1) = 1 and Corollary 1 to derive

V1 =1

2v1 (1, 1) +

1

2v1 (2, 1) ,

where v1 (1, 1) and v1 (2, 1) are given earlier. Since exp (x/λ) is a convex function of x, we obtain

that1

2v1 (1, 1) +

1

2v1 (2, 1) < λ ln

1

2[exp (v1 (1, 1) /λ) + exp (v1 (2, 1) /λ)] .

Since v1 (2, 1) = v1 (1, 2) for the above symmetric interior solution, we deduce that the corner

solution gives a smaller initial value than the above symmetric interior solution. Similarly the

other corner solution in which q1 (2) = 1 is not optimal. Q.E.D.

B Two-Period Case

In this appendix we study problem 3 by dynamic programming. First, consider the problem in

period 2 conditional on the history of a chosen action a1 :

V2 (µ2 (·|a1)) = maxq2(·|a1),µ2(·|·,a1)

∑a2

q2 (a2|a1)Na2H

(µ2

(·|a2))− λH (µ2 (·|a1)) (B.1)

35

Page 36: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

subject to

µ2 (·|a1) =∑a2

µ2

(·|a2)q2 (a2|a1) ,

where the net utility Na2H is given by

Na2H

(µ2

(·|a2))

=∑x2

µ2

(x2|a2

)u (x2, a2) + λH

(µ2

(·|a2)).

It follows from Propositions 1 and 2 that q2(·|a1) and µ2(·|·, a1) are optimal if and only if: (i)

Equation (25) holds. (ii) For any x2 ∈ X, and chosen actions a1 and a2 with q1 (a1) > 0 and

q2 (a2|a1) > 0, we have

V2(x2|a1) = u(x2, a2) + λh′(µ2(x2|a2)) + λf2(µ2

(·|a2)), (B.2)

where

f2(ν) ≡∑x

[h(ν(x))− ν(x)h′(ν(x))

], ν(x) ∈ ∆ (X) . (B.3)

(iii) For any unchosen action a2 and µa22 ∈ ∆ (X) such that

u (x2, a2) + λh′ (µa22 (x))−[u (M,a2) + λh′ (µa22 (M))

]= V2(x2|a1)− V2(M |a1)

for x2 = 1, 2, ...,M − 1, we have

∑x2

I

(V2(x2|a1)

λ− u(x2, a2)

λ− f2(µa22 )

)≤ 1.

Moreover, the value in period 2 satisfies

V2 (µ2 (·|a1)) = V 2(µ2(·|a1))− λH(µ2(·|a1)) =∑x2

µ2 (x2|a1) V2 (x2|a1)− λH (µ2 (·|a1)) . (B.4)

Next consider the problem in period 1. By dynamic programming, we use (B.4) to derive

V1 (µ1) = maxq1,µ1(·|·)

∑a1

q1 (a1)Na1G (µ1 (·|a1))− λH (µ1) , (B.5)

where

µ1(x1) =∑a1

q1(a1)µ1(x1|a1), x1 ∈ X. (B.6)

The net utility Na1G associated with action a1 is

Na1G (µ1(·|a1)) =

∑x1

µ1(x1|a1)u(x1, a1) + βV 2(µ2(·|a1)) + λGa1(µ1(·|a1)), (B.7)

where V 2 is given in (B.4), Ga1 is defined in (31), and µ2 (·|a1) satisfies (25).

36

Page 37: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

It follows from Proposition 1 that V 2 is concave in µ2(·|a1), and hence concave in µ1(·|a1) by

(25). Moreover, Assumption 2 ensures the concavity of Ga1 in µ1(·|a1). Therefore the net utility

Na1G is concave in µ1(·|a1). We view the problem in period 1 as a static RI problem with the prior

belief µ1. Applying Lemma 3 of CD (2013), we derive the following result.

Lemma 2 Let Assumptions 1 and 2 hold. Then the pair µ1(·|·) and q is optimal for problem (B.5)

if any only if it satisfies the following conditions: (i) Equation (24) is satisfied. (ii) There exists a

function V1 (x1) such that for any chosen action a1 with q1(a1) > 0 and for any x1 ∈ X,

V1 (x1) = v1(x1, a1) + λh′(µ1(x1|a1)) + λf1(µ1(·|a1), a1), (B.8)

where

v1(x1, a1) ≡ u(x1, a1) + β∑x2

π(x2|x1, a1)[V (x2|a1)− λh′(µ2(x2|a1))

], (B.9)

f1(ν, a) ≡∑x

[h(ν(x))− ν(x)h′(ν(x))

]− β

∑x′

[h

(∑x

π(x′|x, a

)ν(x)

)−∑x

π(x′|x, a

)ν (x) h′

(∑x

π(x′|x, a

)ν(x)

)].

(B.10)

(iii) For any unchosen action a1 and µa11 ∈ ∆ (X) such that[v1(x1, a1) + λh′(µa11 (x1))

]−[v1(M,a1) + λh′(µa11 (M))

]= V1 (x)− V1 (M) ,

for x1 = 1, 2, ...,M − 1, we have∑x1

I( V1(x1)

λ− v1(x1, a1)

λ− f1(µa11 , a1)

)≤ 1. (B.11)

Moreover, the optimal value for the two-period RI problem satisfies

V1(µ1) =∑x1

µ(x1)V1(x1)− λH(µ1).

Proof: First, we rewrite (26) as

µ2(x2|a1) =M−1∑x1=1

π(x2|x1, a1)µ1(x1|a1) + π(x2|M,a1)

(1−

M−1∑x1=1

µ1(x1|a1)

), (B.12)

and (31) as

Ga1(µ1(·|a1)) =M−1∑x1=1

h(µ1(x1|a1)) + h

(1−

M−1∑x1=1

µ1(x1|a1)

)

− βM−1∑x2=1

h(µ2(x2|a1))− βh

(1−

M−1∑x2=1

µ2(x2|a1)

).

(B.13)

37

Page 38: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Using (B.12) and (B.13), we calculate

∂Ga1(µ1(·|a1))

∂µ1(x1|a1)=h′(µ1(x1|a1))− h′(µ1(M |a1))

− βM∑x2=1

h′(µ2(x2|a1)) [π(x2|x1, a1)− π(x2|M,a1)] ,

(B.14)

for x1 = 1, . . . ,M − 1.

Second, using Proposition 1 (iii) and (B.12), we obtain for x1 = 1, . . . ,M − 1,

∂V 2(µ2(·|a1))

∂µ1(x1|a1)=

M∑x2=1

V2(x2|a1)[π(x2|x1, a1)− π(x2|M,a1)]. (B.15)

Therefore, combining (B.14) and (B.15), we derive

∂Na1G (µ1(·|a1))

∂µ1(x1|a1)=u(x1, a1)− u(M,a1) + β

∂V 2(µ2(·|a1))

∂µ1(x1|a1)+ λ

∂G(µ1(·|a1))

∂µ1(x1|a1)

=[u(x1, a1) + λh′(µ1(x1|a1))]− [u(M,a1) + λh′(µ1(M |a1))]

+ β

M∑x2=1

[V2(x2|a1)− λh′(µ2(x2|a1))

][π(x2|x1, a1)− π(x2|M,a1)]

=[v1(x1, a1) + λh′(µ1(x1|a1))

]−[v1(M,a1) + λh′(µ1(M |a1))

],

(B.16)

where v1 is given in (B.9).

Condition (ED) in Lemma 3 of CD (2013) is equivalent to[v1(x1, a1) + λh′(µ1(x1|a1))

]−[v1(M,a1) + λh′(µ1(M |a1))

]=[v1(x1, b1) + λh′(µ1(x1|b1))

]−[v1(M, b1) + λh′(µ1(M |b1))

],

(B.17)

for any chosen actions a1, b1 and x1 = 1, . . . ,M − 1. Using (B.7) and (B.16), we calculate

Na1G (µ1(·|a1))−

M−1∑x1=1

∂Na1G (µ1(·|a1))

∂µ1(x1|a1)µ1(x1|a1)

=v1(M,a1) + λh′(µ1(M |a1)) + λf1(µ1 (·|a1) , a1),

(B.18)

where f1 is given in (B.10). Then condition (CT) in CD (2013) is equivalent to

v1(M,a1) + λh′(µ1(M |a1)) + λf1(µ1(·|a1), a1) (B.19)

= v1(M, b1) + λh′(µ1(M |b1)) + λf1(µ1 (·|b1) , b1),

for any chosen actions a1 and b1.

The rest of the proof is the same as the proof of Proposition 2 with u, µ(·|·), and f replaced by

v1, µ1(·|·), and f1, respectively. After V1 is identified, the expression for the optimal value follows

from Proposition 1 directly. �

38

Page 39: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

C Forward-backward Algorithm

The algorithm consists of the following steps:

1. Initialize f(0)t (at|at−1) ∈ R, q

(0)t (·|at−1) ∈ ∆ (A) , and µ

(0)t (·|at−1) ∈ ∆ (X) , with q

(0)t (at|at−1) >

0 and µ(0)t (xt|at−1) > 0 for any xt ∈ X, at, at−1 ∈ A, and t = 1, . . . , T . Set µ

(0)1 (·|a0) = µ1 (·).

2. Choose a large integer K. For k = 1, 2, . . . ,K until q(k)t (·|·) gets sufficiently close to q

(k−1)t (·|·)

for all t = 1, . . . , T , do the following:

Backward path: Initialize

v(k)T (xT , aT ) = u(xT , aT ) + β

∑xT+1

π(xT+1|xT , aT )uT+1 (xT+1) .

For t = T, T − 1, . . . , 1 do:

• For each at−1 ∈ A, take v(k)t , f

(k−1)t (·|at−1), q

(k−1)t (·|at−1), and µ

(k−1)t (·|at−1) as the input

u, f(·), q(·), µ for the algorithm in the static case and use equations (47), (49), (50), and

(37) to compute the update f(k)t (·|at−1), q

(k)t (·|at−1), V

(k)t (·|at−1), and µ

(k)t (·|·, at−1).

• If t ≥ 2, compute

v(k)t−1(xt−1, at−1) = u(xt−1, at−1)+β

∑xt

π(xt|xt−1, at−1)[V

(k)t (xt|at−1)−λh′(µ(k−1)

t (xt|at−1))].

Forward path: For t = 1, 2, . . . , T − 1 do:

• If t = 1, compute

µ(k)2 (x2|a1) =

∑x1

π(x2|x1, a1)µ(k)1 (x1|a1).

• If t ≥ 2, compute

µ(k)t+1(xt+1, at) =

∑xt,at−1

π(xt+1|xt, at)µ(k)t (xt|att−1)q

(k)t (at|at−1)µ

(k)t (at−1),

µ(k)t+1(xt+1|at) =

µ(k)t+1(xt+1, at)

µ(k)t+1(at)

, µ(k)t+1(at) =

∑xt+1

µ(k)t+1(xt+1, at) > 0,

where we set µ(k)2 (a1) = q

(k)1 (a1).

3. Return q(K)t (·|·), µ(K)

t (·|·, ·) and V(K)t (·|·).

4. Check whether the converged solution satisfies (48).

For the infinite horizon case, we increase T until convergence.

39

Page 40: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

D Signal-Based Formulation

Following SSM (2017), we consider the signal-based formulation. Suppose that there is a signal

space S satisfying |A| ≤ |S| <∞. At time t, the DM can choose any signal about the state xt with

realizations st in S. A strategy is a pair (d, σ) composed of

1. an information strategy d consisting of a system of signal distributions dt(st|xt, st−1

), for all

st ∈ St, xt ∈ Xt, and t ≥ 1;

2. an action strategy σ consisting of a system of mappings σt : St → A, which give an action

at = σt(st), for t ≥ 1.

Given an action strategy σ, we denote by σt(st)

the history of actions up to time t given the

realized signals st. The state transition kernel π and the strategy (d, σ) induce a sequence of joint

distributions for xt+1 and st recursively

µt+1

(xt+1, st

)= π

(xt+1|xt, σt

(st))dt(st|xt, st−1

)µt(xt, st−1

),

where µ1

(x1, s0

)= µ1 (x1) is given. Using this sequence of distributions, we can compute the

predictive distributions µt(xt|st−1

)and the posteriors µt

(xt|st

)and hence we can define the dis-

counted UPS information cost∑T

t=1 βt−1I

(xt; st|st−1

), similar to (6).

Problem 4 (signal-based dynamic RI problem)

maxd,σ

E

[T∑t=1

βt−1u(xt, σt

(st))

+ βTuT+1 (xT+1)

]− λ

T∑t=1

βt−1I(xt; st|st−1

)where the expectation is taken with respect to the joint distribution over sequences xT+1 and sT

induced by the transition kernel π and the strategy (d, σ) .

We say that a strategy (d, σ) generates a choice rule {pt} if

pt(at|xt, at−1

)= Pr

(σt(st)

= at|xt, σt−1(st−1

)= at−1

)for all at, x

t, and at−1. Conversely, a choice rule {pt} of the form pt(at|xt, at−1

)can induce a

strategy (d, σ) as described by SSM (2017).

We have the following recommendation lemma or the revelation principle similar to Lemma 1

of SSM (2017) or Lemma 2 of Ravid (2019).

Lemma 3 Let Assumption 2 hold. Then any strategy (d, σ) solving the dynamic RI Problem 4

generates sequences of posteriors{µt(xt|at

)}and default rules

{qt(at|at−1

)}solving Problem 1.

Conversely any sequences of posteriors{µt(xt|at

)}and default rules

{qt(at|at−1

)}solving Problem

1 induce a strategy (d, σ) solving Problem 4.

40

Page 41: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Proof: We focus on the finite-horizon case with T < ∞. The result for the infinite-horizon case

can be obtained by taking limits as T →∞.First, using the constructed {pt} from a strategy (d, σ) , we can define a sequence of joint

distributions µt(xt, at−1

)as in (1). The distribution induced by the strategy (d, σ) and the sequence

of distributions µt(xt, at−1

)give the same stream of expected utility. Next we show that the

discounted information cost associated with {pt} ,∑T

t=1 βt−1I(xt; at|at−1), is not larger than that

associated with (d, σ) . These information costs can be computed using the posteriors and predictive

distributions (priors) induced by the corresponding joint distributions.

By the definition of the discounted UPS cost in (6), we compute

I(xt; at|at−1

)=

∑at−1

qt−1

(at−1

)CH

(µt(·|at−1

), µt(·|·, at−1

), qt(·|at−1

))=

∑at−1

qt−1

(at−1

)H(µt(·|at−1

))−∑at−1

qt−1

(at−1

)∑at

qt(at|at−1

)H(µt(·|at))

=∑at−1

qt−1

(at−1

)H(µt(·|at−1

))−∑at

qt(at)H(µt(·|at)).

Rearranging the terms in the discounted UPS cost yields

T∑t=1

βt−1I(xt; at|at−1)

= H(µ1) +T−1∑t=1

∑at

βt−1qt(at)[βH

(µt+1(·|at)

)−H(µt(·|at))

]−βT−1

∑aT

qT (aT )H(µT (·|aT ))

= H(µ1)−T−1∑t=1

∑at

βt−1qt(at)Gat(µt(·|at))− βT−1

∑aT

qT (aT )H(µT (·|aT )). (D.1)

We can derive a similar decomposition for∑T

t=1 βt−1I(xt; st|st−1).

Now we prove that ∑at

qt(at)Gat(µt(·|at)) ≥

∑st

µt(st)Gat(µt(·|st)),

where µt(st) denotes the marginal distribution of st. Since at = σt

(st), we have

µt(xt|at) =∑st

µt(xt|st)Pr(st|at), xt ∈ X.

Since Gat is concave, it follows from Jensen’s inequality that

Gat(µt(·|at)) ≥∑st

Pr(st|at)Gat(µt(·|st)).

41

Page 42: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Multiplying both sides by qt(at) and summing over at, we obtain∑

at

qt(at)Gat(µt(·|at)) ≥

∑st

∑at

Pr(st|at)qt(at)Gat(µt(·|st)) =∑st

µt(st)Gat(µt(·|st)).

Since the generalized entropy H is concave, we can similarly prove that∑aT

qT (aT )H(µT (·|aT )) ≥∑sT

µT (sT )H(µT (·|sT )).

Applying the preceding two inequalities to the second and the third terms on the right-hand

side of (D.1), we obtain

T∑t=1

βt−1I(xt; at|at−1) ≤T∑t=1

βt−1I(xt; st|st−1).

We have shown that the discounted expected payoff from any strategy (d, σ) is not larger than

the value of the objective function in Problem 1 given the sequences of posteriors and default rules

consistent with the joint distribution µT(xT+1, aT

), that is induced by the choice rule generated

by (d, σ). Conversely, using the Bayes rule in (4) to construct the choice rule {pt}, we follow the

same argument as in SSM (2017) to construct a strategy (d, σ) . Notice that the choice rule takes

the form pt(at|xt, at−1

)so that the signal distribution dt depends only on

(xt, s

t−1), but not on

xt−1. The discounted expected payoff from this strategy is identical to the value of the objective

function in Problem 1 given the sequences of posteriors and default rules. These two relationships

together imply the result. �

42

Page 43: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Reference

Arimoto, Suguru, 1972, An Algorithm for Computing the Capacity of Arbitrary Discrete Memo-

ryless Channels, IEEE Transactions on Information Theory 18, 14-20.

Belis, Mariana, and Silviu Guiasu, 1968, A Quantitative and Qualitative Measure of Information

in Cybernetic Systems, IEEE Transactions on Information Theory 14, 593–594.

Blahut, Richard, E., 1972, Computation of Channel Capacity and Rate-Distortion Functions,

IEEE Transactions on Information Theory 18, 460-473.

Caplin, Andrew, and Mark Dean, 2013, Behavioral Implications of Rational Inattention with

Shannon Entropy, NBER working paper 19318.

Caplin, Andrew, and Mark Dean, 2015, Revealed Preference, Rational Inattention, and Costly

Information Acquisition, American Economic Review 105, 2183-2203.

Caplin, Andrew, Mark Dean, and John Leahy, 2019a, Rational Inattention, Optimal Consideration

Sets, and Stochastic Choice, Review of Economic Studies 86, 1061-1094.

Caplin, Andrew, Mark Dean, and John Leahy, 2019b, Rationally Inattentive Behavior: Charac-

terizing and Generalizing Shannon Entropy, working paper, Columbia University.

Cressie, Noel, and Timothy R.C. Read, 1984, Multinomial Goodness-of-Fit Tests, Journal of the

Royal Statistical Society. Series B (Methodological), 46(3), 440–464.

Dewan, Ambuj, and Nathaniel Neligh, 2020, Estimating Information Cost Functions in Models of

Rational Inattention, Journal of Economic Theory 187, 1-32.

Fudenberg, Drew, and Tomasz Strzalecki, 2015, Dynamic Logit with Choice Aversion, Economet-

rica 83, 651-691.

Havrda, Jan, and Frantisek Charvat, 1967, Quantification Method of Classification Processes:

Concept of Structural a-Entropy, Kybernetika 3, 30-35.

Hebert, Benjamin, and Michael Woodford, 2018, Rational Inattention in Continuous Time, work-

ing paper, Columbia University.

Jung, Junehyuk, Jeong-Ho Kim, Filip Matejka, and Christopher A. Sims, 2018, Discrete Actions

in Information-constrained Decision Problems, working paper, Princeton University.

Kamenica, Emir, 2018, Bayesian Persuasion and Information Design, forthcoming in Annual Re-

view of Economics.

43

Page 44: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Luo, Yulei, 2008, Consumption Dynamics under Information Processing Constraints, Review of

Economic Dynamics 11, 366-385.

Mackowiak, Bartosz, and Mirko Wiederholt, 2009, Optimal Sticky Prices under Rational Inatten-

tion, American Economic Review 99, 769–803.

Mackowiak, Bartosz, Filip Matejka, and Mirko Wiederholt, 2018, Survey: Rational Inattention, a

Disciplined Behavioral Model, working paper, Sciences Po.

Matejka, Filip, and Alisdair McKay, 2015, Rational Inattention to Discrete Choices: A New

Foundation for the Multinomial Logit Model, American Economic Review 105, 272-298.

Mattsson, Lars-Goran and Jorgen W. Weibull, 2002, Probabilistic Choice and Procedurally Bounded

Rationality, Games and Economic Behavior 41, 61-78.

Miao, Jianjun, 2019, Multivariate LQG Control under Rational Inattention in Continuous Time,

working paper, Boston University.

Miao, Jianjun, Jieran Wu, and Eric Young, 2019, Multivariate Rational Inattention, working

paper, Boston University.

Mondria, Jordi, 2010, Portfolio Choice, Attention Allocation, and Price Comovement, Journal of

Economic Theory 145, 1837-1864.

Morris, Stephen, and Philipp Strack, 2019, The Wald Problem and the Equivalence of Sequential

Sampling and Static Information Costs, working paper, MIT.

Peng, Lin, and Wei Xiong, 2006, Investor Attention, Overconfidence and Category Learning,

Journal of Financial Economics 80, 563-602.

Puterman, Martin L., 2005, Markov Decision Processes: Discrete Stochastic Dynamic Program-

ming, John Wiley&Sons, New Jersey.

Ravid, Doron, 2019, Bargaining with Rational Inattention, working paper, University of Chicago.

Renyi, Alfred, 1961, On Measures of Entropy and Information, Fourth Berkeley Symposium on

Mathematical Statistics and Probability, 547–561, 1961.

Rust, John, 1987, Optimal Replacement of GMC Bus Engines: An Empirical Model of Harold

Zurcher, Econometrica 55, 999-1033.

Rust, John, 1994, Structural Estimation of Markov Decision Processes, in R. Engle and D. Mc-

Fadden (eds.) Handbook of Econometrics Volume 4, 3082–3139, North Holland.

44

Page 45: Dynamic Discrete Choice under Rational Inattentionpeople.bu.edu/miaoj/UPS04.pdfMa ckowiak and Wiederholt (2009), Mondria (2010), Van Nieuwerburgh and Veldkamp (2010), Miao (2019),

Shannon, Claude, E., 1948, A Mathematical Theory of Communication, Bell System Technical

Journal 27, 379–423.

Sims, Christopher A., 1998, Stickiness, Carnegie-Rochester Conference Series On Public Policy

49, 317–356.

Sims, Christopher A., 2003, Implications of Rational Inattention, Journal of Monetary Economics

50 (3), 665–690.

Sims, Christopher A., 2011, Rational Inattention and Monetary Economics, in Handbook of Mone-

tary Economics, V. 3A, Benjamin M. Friedman and Michael Woodford (ed.), North-Holland.

Steiner, Jakub, Colin Stewart, and Filip Matejka, 2017, Rational Inattention Dynamics: Inertia

and Delay in Decision-Making, Econometrica 85, 521-553.

Stokey, Nancy L., Robert E. Lucas with Edward C. Prescott, 1989, Recursive Methods in Economic

Dynamics, Harvard University Press.

Tanaka, Takashi, Henrik Sandberg, Mikael Skoglund, 2018, Transfer-Entropy-Regularized Markov

Decision Processes, working paper.

Tsallis, Constantino, 1988, Possible Generalization of Boltzmann-Gibbs Statistics, Journal of

Statistical Physics 52, 479–487.

Van Nieuwerburgh, Stijn, and Laura Veldkamp, 2010, Information Acquisition and Under-diversification,

Review of Economic Studies 77, 571-626.

Woodford, Michael, 2009, Information Constrained State Dependent Pricing. Journal of Monetary

Economics 56(S), S100–S124.

Woodford, Michael, 2012, Inattentive Valuation and Reference-dependent Choice, working paper,

Columbia University.

Zhong, Weijie, 2019, Optimal Dynamic Information Acquisition, working paper, Columbia Uni-

versity.

45