Optimal Policy Rules in HANK
Alisdair McKay Christian K. Wolf
FRB Minneapolis MIT & NBER
March 13, 2023
Abstract: We characterize optimal policy rules in business-cycle models with
nominal rigidities and heterogeneous households. The derived rules are expressed
in terms of the causal effects of policy instruments on policymaker targets. Our
first result is that the optimal policy rule of a “dual mandate” central banker—a
policymaker that only cares about inflation and output—is unaffected by house-
hold heterogeneity. The optimal rule of a Ramsey planner contains an additional
distributional term that incorporates the effects of her available policy instru-
ments on consumption inequality. We quantify the role of this term in a structural
model that incorporates and is consistent with evidence on the key distributional
channels of monetary policy, and we find that the concern for inequality only
has a moderate effect on optimal interest rate policy. Fiscal stimulus payments,
on the other hand, have strongly progressive effects and are thus well-suited to
cushion the distributional effects of cyclical fluctuations.
Email: [email protected] and [email protected]. We thank Adrien Auclert, Anmol Bhan-
dari, Jordi Gal´ı,
´
Emilien Gouin-Bonenfant, and Ludwig Straub for helpful comments. The views expressed
herein are those of the authors and not necessarily those of the Federal Reserve Bank of Minneapolis or the
Federal Reserve System.
1
1 Introduction
Should household inequality affect the conduct of cyclical stabilization policy? Recent years
have seen a surge of interest in this question, with several prominent studies arguing that the
design of optimal monetary policy is substantially altered by distributional considerations
(e.g., Bhandari et al., 2021; Acharya et al., 2022). In principle, household heterogeneity can
affect optimal policy design in two separate ways. First, household inequality could alter the
transmission from policy instruments to any given target (e.g., inflation and output). This
change in transmission may affect whether or not a policymaker can attain her given targets,
and how policy instruments need to be set to do so. Second, household heterogeneity may
also alter the targets themselves. For example, with market incompleteness, policymakers
may want to dampen the distributional effects of cyclical fluctuations.
We provide new insights into this question by expressing optimal policy rules in business-
cycle models with household heterogeneity in terms of empirically measurable “sufficient
statistics.” Our first contribution is to cast optimal policy problems in a rich heterogeneous-
agent New Keynesian (HANK) environment in linear-quadratic form, closely mirroring the
canonical representative-agent New Keynesian (RANK) literature. As familiar from this lit-
erature, the solution to the linear-quadratic policy problem takes the form of a forecast target
criterion, providing a general characterization of optimal policy completely independently of
the shocks hitting the economy (Giannoni & Woodford, 2002). Drawing on McKay & Wolf
(2023), we furthermore show that these optimal policy rules can be expressed in terms of the
dynamic causal effects of policy instruments on policymaker targets. This “sufficient statis-
tics” characterization of optimal policy rules has two benefits. First, it provides a convenient
organizing device for understanding the literature—differences in optimal policy design can
be traced back to differences in the causal effects of the policy instrument on policymaker
targets. Second, it suggests a natural avenue for model discipline—connect as tightly as
possible with existing empirical evidence on the causal effects of policy interventions.
We then use our sufficient statistics framework to revisit the two ways in which household
heterogeneity can in principle affect optimal policy design. On transmission (i.e., fixing
policymaker targets), we show that the optimal policy rule of a dual mandate central banker
does not change with household heterogeneity. Such heterogeneity may affect the path of
interest rates required to implement the dual mandate optimum, but our sufficient statistics
perspective reveals that the scope even for this is rather limited. On targets, in principle,
distributional concerns can have large effects on optimal stabilization policy. In practice,
2
however, empirical evidence suggests that monetary policy has broad-based effects in the
cross-section of households, with consumption of all households rising following a monetary
easing, and vice-versa. Monetary policy is thus a rather blunt distributional tool, and as a
result a Ramsey planner in an environment consistent with such broad-based effects does
not choose to deviate much from dual mandate policy prescriptions.
Environment. Our analysis is set in a business-cycle model with nominal rigidities and
household heterogeneity. Households face idiosyncratic income risk and self-insure by bor-
rowing and saving in capital, short-term bonds, and long-term bonds. The policymaker sets
short-term nominal interest rates, pays transfers to households, and finances its expenditure
through taxation as well as bond issuance. As we discuss later, the environment is designed
to be rich enough to speak to the core channels of how monetary policy affects consumption
across the cross-section of households.
In this environment we study optimal policy problems cast in linear-quadratic form (e.g.,
Giannoni & Woodford, 2002; Benigno & Woodford, 2012). Linearization of the private-
sector relations of our economy yields the linear constraints. Our quadratic objective is then
either ad hoc—a conventional dual mandate objective—or derived as an approximation to a
particular Ramsey policy problem.
Dual-mandate central banker. We begin our analysis by studying the optimal policy
problem of a conventional “dual-mandate” central banker that seeks to close the output gap
and stabilize inflation. We find this ad hoc loss function useful as a starting point, for two
reasons. First, conceptually, it allows us to explore the effects of changes in policy instrument
propagation while fixing policymaker targets. Second, practically, it is arguably the relevant
objective function for real-world central banks.
Our main result is that the optimal interest rate target criterion for such a dual mandate
central banker is exactly the same as in a standard representative-agent environment. The
logic is as follows. Household heterogeneity in our environment only affects the demand side
of the economy (i.e., the model’s “IS” curve). In the optimal policy problem, however, this
demand block is a slack constraint: the policymaker can pick an output-inflation allocation
subject to the Phillips curve, and then set nominal interest rates as necessary to generate
demand consistent with the desired allocation. Household heterogeneity thus may matter
for the path of the policy instrument, but not for optimal output and inflation outcomes.
Leveraging our sufficient statistics perspective, we then connect these theoretical results
on optimal dual-mandate policy to empirical evidence on monetary policy shock transmis-
3
sion. The empirical literature has tried to identify the causal effects of nominal rate changes
on output and inflation (e.g. see Ramey, 2016). While silent on transmission mechanisms
and so in particular on the importance of heterogeneity-related channels, aggregate data are
informative about the size of interest rate movements required to move output and infla-
tion by a given amount. But it then follows that quantitatively relevant HANK and RANK
models—that is, models that are consistent with the empirical evidence on monetary shock
propagation—will not only share the same optimal dual-mandate policy rule, but they will
in fact also tend to agree quite closely on the interest rate path needed to implement the
optimal inflation and output outcomes. Our quantitative analysis confirms this insight: our
HANK model—which is calibrated to agree with this empirical evidence on monetary shock
propagation—agrees closely with a RANK model calibrated to the same evidence.
Ramsey problem. We next study a Ramsey problem in which the planner seeks to max-
imize a weighted sum of individual household utilities, allowing us to explore the role of
changes in policymaker objectives. Applying a second order approximation, we derive a loss
function that adds to the usual output gap and inflation objectives a third term reflect-
ing distributional concerns. We then solve the policymaker problem of minimizing this loss
function subject to linearized private-sector optimality conditions.
The policymaker’s problem now yields an optimal policy rule consisting of three terms,
trading off the policymaker’s ability to use her policy instrument to stabilize output, infla-
tion, and the cross-sectional consumption distribution. The third term of this rule—which,
importantly, is the sole difference from the optimal standard dual mandate policy rule—is
governed by the distributional incidence of the instrument. If, for example, interest rate
movements do not affect consumption shares (e.g., as in Werning, 2015), then the third term
is zero and the Ramsey rule is identical to the optimal dual-mandate rule; if, on the other
hand, monetary policy has large distributional effects (e.g., as in Bhandari et al., 2021), then
distributional concerns may swamp price and output stability considerations.
To characterize the quantitatively relevant Ramsey policy, we again leverage our sufficient
statistics perspective. As we have discussed, our conclusions on optimal policy will be
governed by the causal effects of interest rate policy on consumption inequality. Since prior
empirical work has delivered the sharpest results on the cross-sectional income incidence of
monetary policy, our strategy is to calibrate our model to match these distributional channels
of monetary policy, and then rely on the structure of the model to infer the ultimate effects on
consumption inequality. Two features of our model turn out to be particularly relevant. First,
4
our model features long-duration assets, allowing us to capture an important redistribution
channel of monetary policy (see Auclert, 2019). Second, the model is designed to generate
only a modest response of aggregate labor income following a monetary expansion, consistent
with empirical evidence. For our purposes, the key takeaway is that the model generates
fairly evenly distributed effects of policy, with consumption responding by similar percentage
amounts across the wealth and income distribution. Given this particular incidence, we find
that monetary policy is rather ill-suited as a tool to deal with the distributional implications
of business-cycle shocks. For example, if—in response to a shock that largely affects the
consumption of the poor—interest rates were set to stabilize consumption at the bottom,
then consumption of other households and so aggregate output and inflation would overshoot.
The Ramsey policymaker does not find such overshooting optimal, and instead responds to
the shock similarly to conventional optimal dual-mandate policy. These quantitative findings
contrast with recent work on optimal monetary policy with heterogeneous households that
tends to find an important role for distributional considerations (e.g., Bhandari et al., 2021;
Acharya et al., 2022; avila & Schaab, 2022).
Finally, we also briefly consider a Ramsey policymaker that jointly sets interest rates and
stimulus checks, consistent with recent U.S. policy practice. While equivalent in their effects
on output and inflation (Wolf, 2021), the two instruments differ significantly in their distribu-
tional incidence, with fiscal stimulus payments sharply compressing consumption inequality.
The two instruments are thus highly complementary, with monetary policy primarily aimed
at aggregate stabilization while fiscal policy manages distributional concerns.
Literature. We contribute to the literature on optimal policy in business-cycle models
with rich microeconomic heterogeneity (Acharya et al., 2022; Bhandari et al., 2021; Le Grand
et al., 2022; avila & Schaab, 2022). Conceptually, our key contribution is to characterize
optimal policy through forecast target criteria expressed in terms of easily interpretable suf-
ficient statistics. Our computation of these optimal policy rules leverages sequence-space
representations of equilibria and thus the recent work of Auclert et al. (2021).
1
A con-
temporaneous paper that does the same is avila & Schaab (2022). Those authors do not
rely on linear-quadratic approximations, thus providing more general optimal policy results,
but without our sufficient statistics perspective. Our discussion of the mapping between
1
By the equivalence of perfect-foresight sequence-space and stochastic linear state-space methods, our tar-
geting criterion also applies to the analogous stochastic linear-quadratic optimal control problem. Sequence-
space linear-quadratic policy problems have been used in prior work to derive “optimal policy projections”
(Svensson, 2005; De Groot et al., 2021; Hebden & Winker, 2021).
5
policy rules and empirical evidence builds on our earlier work in McKay & Wolf (2023).
These conceptual innovations allow us to revisit several substantive results in the prior
heterogeneous-household optimal policy literature.
First, we derive a sharp set of irrelevance results for the effects of inequality on optimal
policy design. Prior work has emphasized that inequality will affect policy propagation by
altering the demand side of the economy (Kaplan et al., 2018; Auclert, 2019); our analysis,
however, reveals that these changes do not affect the optimal “dual mandate” output and in-
flation outcomes. If furthermore matched to the same evidence on policy shock propagation,
then HANK and RANK models will also (roughly) agree on the required rate paths.
Second, our sufficient statistics perspective provides novel insights on the insurance role
of stabilization policy. Our formulation of the Ramsey problem objective isolates the pol-
icymaker’s desire to provide insurance against aggregate shocks with unequal incidence on
consumption across households.
2
In particular, our second-order approximation to the social
welfare function extends a similar finding by Acharya et al. to a more traditional HANK
model that does not permit analytical aggregation. The key benefit of our formulation is that
it delivers a transparent characterization of optimal policy, with the extent to which house-
hold heterogeneity affects the optimal Ramsey policy fully governed by the distributional
effects of the policy instrument. Guided by these insights, and differently from prior work
(e.g., Bhandari et al., 2021; Le Grand et al., 2022), we consider a rich quantitative model
that is consistent with the main distributional channels of monetary stabilization policy.
Finally, our analysis of optimal joint fiscal-monetary policy extends results in Wolf (2021)
and Bilbiie et al. (2021). Wolf (2021) shows that, in a standard HANK setting, transfers and
monetary policy can implement the same aggregate allocations, but differ in their distribu-
tional implications. In a two-agent environment, Bilbiie et al. (2021) argue that monetary
and fiscal policy can together stabilize both the aggregate activity as well as the consump-
tion shares of the two household types. Our work is complementary: we analyze the optimal
monetary-fiscal policy mix in an environment with rich heterogeneity.
Outline. In Section 2, we present a general linear-quadratic sequence-space problem and
its solution. Section 3 outlines our HANK model, and Sections 4 and 5 discuss our results
for optimal dual-mandate and Ramsey policy, respectively. We conclude in Section 6.
2
As such, our formulation of the problem abstracts from many of the issues of time consistency and
inflation bias discussed in avila & Schaab (2022).
6
2 Linear-quadratic problems in the sequence space
Throughout this paper we study optimal policy problems that can be recast as deterministic
linear-quadratic control problems. We begin in Section 2.1 by first stating the problem and
presenting its solution. Section 2.2 then places our analysis in the broader context of the
literature, focussing in particular on the tight connection between our expressions and the
(in principle estimable) causal effects of policy shocks.
2.1 Problem & solution
We consider a policymaker that faces a linear-quadratic optimal control problem.
Preferences. The policymaker targets I variables, indexed by i. We let x
it
be the
deviation of the ith target variable from its target value at date t. We consider a policymaker
with quadratic loss function
L
1
2
X
t=0
β
t
I
X
i=1
λ
i
x
2
it
=
1
2
x
x
x
0
W )x
x
x, (1)
where x
x
x
i
(x
i0
, x
i1
, . . . )
0
is the perfect-foresight sequence of the ith target variable through
time and x
x
x (x
x
x
1
,x
x
x
2
, . . . ,x
x
x
I
)
0
stacks those paths for all of the I targets. The λ
i
’s denote the
weights associated with the different policy targets, with Λ diag(λ
1
, λ
2
, . . . , λ
I
). Finally
W = diag(1, β, β
2
, . . . ) summarizes the effects of discounting in the policymaker preferences,
with discount factor β (0, 1).
Constraints. The policymaker faces constraints imposed by the equilibrium relationships
between variables. These linear constraints are expressed compactly as
H
x
x
x
x + H
z
z
z
z + H
ε
ε
ε
ε = 0
0
0, (2)
where z
z
z (z
z
z
1
,z
z
z
2
, . . . ,z
z
z
J
)
0
stacks time paths for the J policy instruments available to the
policymaker, and ε
ε
ε (ε
ε
ε
1
,ε
ε
ε
2
, . . . ,ε
ε
ε
Q
)
0
similarly stacks the paths for Q exogenous shocks.
{H
x
, H
z
, H
ε
} are then conformable linear maps.
While the structural models considered in the remainder of this paper directly map into
constraints of the general form (2), it follows from the discussion in McKay & Wolf (2023)
7
that these constraints can equivalently and more conveniently be expressed as
x
x
x
i
=
J
X
j=1
Θ
x
i
,z
j
z
z
z
j
+
Q
X
q=1
Θ
x
i
q
ε
ε
ε
q
, i = 1, 2, . . . , I (3)
where the Θ’s are linear maps that capture the dynamic causal effects of a policy instrument
path z
z
z
j
or shock path ε
ε
ε
q
on a target variable path x
x
x
i
. The alternative constraint (3) thus
expresses the policy targets directly in terms of impulse responses to policy instruments and
exogenous shocks, as opposed to imposing implicit relationships as in (2).
3
Problem & solution. The optimal policy problem is to choose the instrument paths
z
z
z to minimize (1) subject either to (2) (for the original constraint formulation) or (3) (for
the re-cast constraint). The policymaker thus minimizes a convex objective subject to linear
constraints, and so the first-order conditions are necessary and sufficient for a solution to
the optimal policy problem.
For intuition, it is simpler and more instructive to use the constraint (3). Minimizing (1)
subject to (3) yields:
1. Optimal policy rule. For each policy instrument z
z
z
j
, the paths of the policy targets
satisfy the “policy criterion”
I
X
i=1
λ
i
· Θ
0
x
i
,z
j
W
|
{z }
(discounted) causal effect of z
j
on x
i
· x
x
x
i
= 0
0
0, j = 1, 2, . . . , J (4)
(4) is simply the first-order condition of the optimal policy problem. It says that, for each
instrument z
z
z
j
, the paths of the policy targets x
x
x
i
must be at an optimum within the space
implementable through z
z
z
j
. In the language of Svensson (1997) and Woodford (2003), this
rule is an example of a so-called implicit “target policy criterion”: the policymaker sets the
available instruments to align projections (i.e., future paths) of macro aggregates as well
as possible with its targets, given what is achievable through the available instruments.
3
The equivalence of (2) and (3) would be immediate for invertible H
x
. In typical macroeconomic models,
however, H
x
is not invertible, so recasting the constraint as (3) requires additional arguments. McKay & Wolf
(2023) provide those arguments; briefly, the core intuition is that the optimal policy problem can be shown
to be equivalent to the alternative, artificial problem of picking shocks to a given baseline, determinacy-
inducing policy rule. Policy and non-policy shocks relative to this arbitrary baseline policy rule then yield
the impulse response matrices Θ. See Appendix C.1 for further details.
8
We emphasize two important features of such rules. First, they are derived without
reference to and so apply independently of the non-policy shocks hitting the economy.
This robustness property is one of the main virtues of target policy criteria (Giannoni &
Woodford, 2002). Second, note that the optimal policy rule for instrument j places no
weight on a policy target i that cannot be moved by instrument j (i.e., Θ
x
i
,z
j
= 0), even
if λ
i
> 0—intuitively, if an instrument does not affect a target, then this target plays no
role in setting the instrument.
2. Optimal policy path. Given the exogenous shock paths ε
ε
ε, the policy rule (4) together
with the constraints (3) characterizes the evolution of the dynamic system. In particular,
the optimal instrument path z
z
z
satisfies
z
z
z
Θ
0
x,z
W
x,z
1
×
Θ
0
x,z
W
x,ε
ε
ε
ε
, (5)
where Θ
x,z
and Θ
x,ε
suitably stack the individual Θ
x
i
,z
j
’s and Θ
x
i
q
’s. The optimal path
of the policy instruments thus has an intuitive regression interpretation: the instruments
z
z
z are set to offset as well as possible—in a weighted least-squares sense—the perturbation
to the policy targets x
x
x caused by the exogenous shocks, given as Θ
x,ε
× ε
ε
ε. In particular,
the policymaker will rely most heavily on the tools z
j
that are best suited to offset the
perturbation to its targets induced by a particular shock path ε
ε
ε.
2.2 Discussion
Equations (4) and (5) will guide our analysis in much of the remainder of the paper. In
this section we briefly relate our results to prior work on: first, stochastic linear-quadratic
problems; and second, empirical measurement of the propagation of policy shocks.
Deterministic transitions vs. aggregate risk. It is well-established that, by cer-
tainty equivalence, the first-order perturbation solution of models with aggregate risk is
mathematically identical to linearized perfect-foresight transition paths.
4
This insight im-
plies the following connections between our linear-quadratic perfect foresight problem and
the canonical linear-quadratic stochastic problem (as in Benigno & Woodford, 2012). First,
the policy target criterion (4) corresponds to a forecast targeting criterion in a stochastic
4
For detailed discussions of this point see for example Fern´andez-Villaverde et al. (2016), Boppart et al.
(2018) or Auclert et al. (2021).
9
economy. For a time-0 problem with commitment, that forecast targeting criterion is simply
E
0
"
I
X
i=1
λ
i
· Θ
0
x
i
,z
j
W · x
x
x
i
#
= 0
0
0, j = 1, 2, . . . , J (6)
This is an implicit rule that determines the expected evolution of the economy as of date 0. In
a stochastic environment, new shocks will occur as time goes by, causing the evolution of the
economy to deviate from what was expected at date 0. In this case, (6) gives a rule for how to
revise forecasts at each date. Second, by the same logic, the optimal instrument path z
z
z
in (5)
corresponds to the instrument impulse response to a time-0 shock that changes expectations
of the exogenous shifters from 0
0
0 to ε
ε
ε. The exact same impulse response interpretation applies
to our solution for the paths of the policy targets x
x
x
.
Measurement and interpreation. The matrices stacked in Θ
x,z
collect the dynamic
causal effects of variations in the policy instruments z onto the policymaker targets x. In
other words, the effects of policy instruments on policymaker targets are sufficient statistics
for the characterization of optimal policy rules. This observation will guide much of our
analysis Sections 4.3 and 5.4, and in particular will allow us to transparently relate our find-
ings to those reported in previous work. We also note a useful connection to measurement:
as we show formally in McKay & Wolf (2023), entries of the causal effect maps Θ
x,z
can be
estimated using semi-structural time series methods applied to identified policy shocks (e.g.,
as in Ramey, 2016). We will thus in the subsequent analysis relate our findings as much as
possible to empirical evidence on the effects of (monetary) stabilization policy.
Outlook. In the remainder of this paper we will first show that optimal policy problems in
models with household heterogeneity can be represented in the form of our linear-quadratic
control problem, and then use (4) and (5) to characterize optimal policy rules. Section 3
begins the analysis with a description of our model environment.
3 Model
We consider a HANK economy with two special features. First, labor supply is intermediated
by labor unions. This will allow us to summarize the supply block through a standard Phillips
curve and thus cleanly focus on the demand-side implications of household heterogeneity.
Second, households invest in a menu of assets, which allows us to capture several empirically
10
relevant channels of redistribution induced by monetary policy. Previewing our applications
in Sections 4.3 and 5.4, we note that the model is designed to be just rich enough to be able
to speak to evidence on the aggregate and distributional implications of monetary policy.
Time is discrete and runs forever, t = 0, 1, 2, . . . . Consistent with our linear-quadratic
framework in Section 2, we will consider linearized perfect-foresight transition sequences.
By certainty equivalence, our solutions will be identical to the analogous economy with
aggregate risk and solved using conventional first-order perturbation techniques with respect
to aggregate variables. Throughout this section, boldface denotes time paths (so e.g., x
x
x
(x
0
, x
1
, x
2
, . . . )
0
), bars indicate the model’s deterministic steady state (¯x), and hats denote
(log-)deviations from the steady state (bx).
5
3.1 Households
The economy is populated by a unit continuum of ex-ante identical households indexed by
i [0, 1]. Household preferences are given by
E
0
X
t=0
β
t
c
1γ
it
1
1 γ
ν (`
it
)
, (7)
where c
it
is the consumption of household i and `
it
is its labor supply.
Households face uninsurable risk to their individual incomes. Let ζ
it
be an idiosyncratic
stochastic event that determines the idiosyncratic component of household i’s income at date
t. The idiosyncratic event ζ
it
follows a stationary Markov process. Following Werning (2015)
and Alves et al. (2020), we assume there is an incidence function Φ that maps aggregate
labor income to individual labor income as a function of ζ
it
. Letting e
it
be the labor earnings
of the household, we have
e
it
= Φ(ζ
it
, m
t
, (1 α)y
t
),
where m
t
is a distributional shock that tilts the incidence function towards or away from
high-income households, and (1 α)y
t
is aggregate labor income. As we describe in greater
detail below, labor as a whole receives a share 1α of total income y
t
. The incidence function
Φ thus satisfies
R
e
it
di = (1 α)y
t
for any value of m
t
and y
t
, and so the shock m
t
only
affects the distribution of labor income, not the total amount. For the quantitative analysis
in Sections 4.3 and 5.4, the shock m
t
will be our example of an inequality shock—a shock
5
To be precise, we use log deviations for the variables {y, c, `, w, b, a, 1+r, 1+i, q
b
, q
k
, η} and level deviations
for the variables {π, τ
x
, τ
e
, m}.
11
that affects aggregate demand through redistribution and precautionary savings motives.
Next, labor supply is determined by a labor market union (described further below), so
hours worked `
it
are taken as given by the household. Total labor income is taxed at some
constant proportional rate τ
y
. Finally households also receive a time-varying lump-sum
transfer τ
x,t
+ τ
e,t
e
it
. The first component of the transfer, τ
x,t
, is the same for all households
and will be manipulated as part of the optimal policy problem; we will refer to it as a “fiscal
stimulus payment” as it resembles the real-world stimulus checks that have been used in
recent recessions in the U.S. The second component, τ
e,t
e
it
, is the “endogenous” component,
adjusting slowly over time to maintain long-run budget balance. This component of transfers
is proportional to the household’s productivity.
Households can save and possibly borrow in a variety of assets with different durations
and different exposures to surprise inflation. Due to certainty equivalence and no-arbitrage,
the returns on all of these assets must be equal at all dates along the equilibrium transition
path except possibly at t = 0, where revaluation effects can lead to heterogeneous realized
returns across households. Let r
t
denote the (equalized) return between t and t + 1, and
furthermore let a
it
denote the net worth of household i at the beginning of period t (inclusive
of interest). The overall budget constraint of household i is then
1
1 + r
t
a
it+1
+ c
it
= a
it
+ (1 τ
y
+ τ
e,t
)e
it
+ τ
x,t
. (8)
All date-0 revaluation effects will be captured by the initial asset position a
i0
. We will discuss
these revaluation effects later. For now, we note that the distribution of households over
initial states (ζ
i0
, a
i0
) is endogenous, and we write it as Ψ
0
. Finally, we impose a constraint
on total household net worth: a
it+1
a, where a 0.
The solution to each individual household i’s consumption-savings problem gives a map-
ping from paths of aggregate income y
y
y, real returns r
r
r, transfers τ
τ
τ
x
and τ
τ
τ
e
, shocks m
m
m, and the
initial asset position a
i0
to the path of consumption c
c
c
i
. Aggregating consumption decisions
across all households, we thus obtain an aggregate consumption function C():
c
c
c = C(y
y
y, r
r
r,τ
τ
τ
x
,τ
τ
τ
e
,m
m
m; Ψ
0
). (9)
3.2 Technology, unions, and firms
Final goods in our economy are produced by a representative firm that combines intermediate
inputs. Those intermediate goods in turn are produced out of capital and labor, with labor
12
supply intermediated by a union and the total capital stock of the economy fixed—an extreme
form of capital adjustment costs.
Final goods production. The production function of the final goods producer is
y
t
=
Z
j
y
η
t
1
η
t
jt
dj
η
t
η
t
1
.
The elasticity of substitution between different intermediate inputs, η
t
, is allowed to vary
(exogenously) over time. This “aggregate supply”-type shock will be important in our dis-
cussion of optimal policy. The final good is sold at nominal price p
t
.
Factor supply. A labor market union intermediates household labor supply. We assume
that hours worked are rationed equally across households, so `
it
= `
t
for all i. Given separable
preferences and with everyone supplying an equal amount of hours worked, it follows that all
households share a common marginal disutility of labor. The marginal utility of consumption,
however, is generally not equalized. For reasons that we will discuss in detail later, we assume
that the union evaluates the benefits of higher after-tax income using the marginal utility of
average consumption (c
γ
t
) rather than the average of marginal utilities (
R
1
0
c
γ
it
di), as also
done in Wolf (2021) and Auclert et al. (2021). Under those assumptions, the labor union
agrees to supply `
t
total units of labor at real wage
(1 τ
y
)w
t
=
ν
0
(`
t
)
c
γ
t
. (10)
The aggregate capital stock is fixed at k. We assume that capital owners must pay p
t
δ
units of the final good in order to maintain each unit of the capital stock.
Intermediate goods producers. A unit continuum of intermediate goods producers
combine capital and labor according to the production technology
y
jt
= Ak
α
jt
`
1α
jt
. (11)
Let w
t
be the real wage and let r
k
t
be the real rental rate for capital. Cost minimization by
the intermediate producers implies they all use the same capital-labor ratio, with
`
jt
=
r
k
t
w
t
1 α
α
k
jt
13
Integrating across firms and using input market clearing we have
w
t
`
t
1 α
=
r
k
t
k
t
α
.
It thus follows that labor receives a share 1 α of factor payments, while capital receives
the residual share α. Real marginal cost is common across firms and equal to
µ
t
=
1
A
r
k
t
α
α
w
t
1 α
1α
.
Total income is y
t
, with a share µ
t
of that total income going to factor payments and the
residual share 1 µ
t
going to monopoly profits. We then furthermore assume that a share
1 α of the monopoly profits is paid to workers in the form of profit sharing, which implies
that labor receives a share 1 α of total aggregate income y
t
(see Appendix A.1). The union
then distributes this total labor income to households according to the incidence function Φ
introduced above. The remaining share of profits is distributed to capital owners, so capital
owners receive a share α of aggregate income.
Our assumption on profit sharing fixes the labor share at 1α. In making this assumption
we have removed one potential source of redistribution from policy. The alternative, and
seemingly more natural, assumption that all monopoly profits are paid to capital owners
implies a sharp rise in the labor share following expansionary monetary policy. However, as
we discuss further below, empirical evidence suggests that monetary expansions have little
effect on the labor share (or perhaps even lower it), and so we have designed our model to
be consistent with that evidence.
Each intermediate goods firm sets its price in standard Calvo fashion, with probability
1 θ of updating the price each period. We show in Appendix A.1 that the solution to this
problem gives rise to a standard linearized perfect-foresight New Keynesian Phillips curve:
bπ
t
= κby
t
+ βbπ
t+1
+ ψbη
t
, (12)
where the composite parameters κ and ψ are functions of model primitives and defined in
Appendix A.1. We furthermore allow for a (time-invariant) subsidy on production financed
with lump-sum taxes on the intermediate goods producers; this subsidy will matter in Sec-
tion 5, where we require efficiency of the deterministic steady state to write our optimal
policy problem in a form consistent with the linear-quadratic set-up of Section 2, exactly as
in prior work (e.g. Woodford, 2003).
14
Aggregate production function. Combining the production functions of final goods
producers and intermediate goods producers, we can write aggregate production as
y
t
=
Ak
α
d
t
`
1α
t
, (13)
where d
t
1 captures efficiency losses from price dispersion.
3.3 Asset structure
There are three different assets: a short-term nominal bond, a long-term nominal bond, and
capital. As discussed above, by no-arbitrage, all assets will provide the same returns along
any equilibrium transition path except possibly at t = 0. This section presents the date-t 1
no-arbitrage relations as well as the date-0 revaluation effects.
Asset returns. The three assets pay out the following real returns.
1. Short-term bond. The short-term nominal interest rate is denoted i
t
. If p
t
dollars are
invested in the short-term bond at date t, then the payoff is valued at (1 + i
t
)p
t
/p
t+1
=
(1 + i
t
)/(1 + π
t+1
) units of goods in the next period.
2. Long-term bond. Households can purchase a unit of the long-term bond for a real price
of q
b
t
. At time t + 1, the household receives a real coupon of (¯r + σ
b
)(1 + π
t+1
)
1
and
furthermore retains a fraction (1 σ
b
)(1 + π
t+1
)
1
of the asset position, now valued at
(1σ
b
)(1+π
t+1
)
1
q
b
t+1
. The parameter σ
b
controls the maturity of the asset, with coupons
decaying at rate σ
b
, while the coupon scaling factor (¯r + σ
b
) normalizes the steady-state
price of the bond to one. The inflation term captures the fact that inflation reduces the
real value of all future nominal payouts.
3. Capital. A unit of capital can be purchased for q
k
t
units of the final good. It follows from
our discussion in Section 3.2 that the real payoff of a unit of capital is r
k
t+1
δ + (1
µ
t+1
)αy
t+1
/k + q
k
t+1
= αy
t+1
/k δ + q
k
t+1
.
Arbitrage relations. By no-arbitrage, all assets must yield the same expected return
at all dates t 1. Letting r
t
denote the common return on these assets for t 1, it then
follows from the above discussion that in equilibrium we must have
1 + r
t
=
1 + i
t
1 + π
t+1
(14)
15
1 + r
t
=
(¯r + σ
b
)(1 + π
t+1
)
1
+ (1 σ
b
)(1 + π
t+1
)
1
q
b
t+1
q
b
t
(15)
1 + r
t
=
αy
t+1
/k δ + q
k
t+1
q
k
t
(16)
Revaluation effects. At time-0, the returns are not necessarily equalized, reflecting
the arrival of surprise shocks. The expressions on the right-hand sides of (14), (15), and (16)
then give us realized returns at date-0 with i
1
, q
b
1
, and q
k
1
at their steady state values.
At the end of period t = 1, the households in our economy had positions in short-
term bonds, long-term bonds, and equities—positions that are then revalued as the date-0
news arrives. We take this initial distribution of portfolios as given; in fact, for our later
quantitative analysis, we will match it directly to empirical evidence on household portfolios.
It follows from the above discussion of revaluation effects that the date-0 distribution of
household asset holdings inclusive of returns, Ψ
0
, depends on π
0
, q
b
0
, q
k
0
, and y
0
. We will
write this mapping as
Ψ
0
= H(y
0
, π
0
, q
b
0
, q
k
0
). (17)
Note that we can use (17) to substitute out for Ψ
0
in the consumption function (9). Lin-
earizing around the deterministic steady state yields the aggregate consumption function
b
c
c
c = C
y
b
y
y
y + C
r
b
r
r
r + C
x
b
τ
τ
τ
x
+ C
e
b
τ
τ
τ
e
+ C
m
b
m
m
m + C
Ψ
H
y
by
0
+ H
π
bπ
0
+ H
b
bq
0
b
+ H
k
bq
0
k
, (18)
where all of the derivative matrices C
and H
are evaluated at the economy’s deterministic
steady state.
3.4 Government
The final actor in our model is the government. The government collects tax revenue, pays
out lump-sum transfers, sets the nominal interest rate on the short-term bond, and issues
short- as well as long-term bonds. Letting a
g
t
denote the value of claims on the government
entering period t (inclusive of returns), the government budget constraint becomes
a
g
t+1
1 + r
t
= a
g
t
+ τ
x,t
(τ
y
τ
e,t
) (1 α)y
t
| {z }
tax revenue
. (19)
Government debt maturity. We assume that the government issues both long-term as
well as short-term bonds. For all dates t 1 along the perfect foresight transition path, the
16
returns on the two bonds are equalized and so the maturity structure of the government debt
is irrelevant. At date 0, on the other hand, the maturity structure matters via revaluation,
perfectly analogously to our earlier discussion of valuation effects in household portfolios. In
particular, the revaluation effects will again be embedded in the date-0 portfolio value a
g
0
.
The aggregate bond holdings of the household sector are the liabilities of the government, and
so the revaluation of the government debt is necessarily the mirror image of the revaluation
of the household bond positions.
Policy instruments. We consider the nominal rate of interest i
t
and the exogenous
component of transfers τ
x,t
(i.e., “stimulus checks”) as the independent policy instruments of
the government, used for business-cycle stabilization policy. We assume that the endogenous
component of transfers τ
e,t
adjusts gradually to ensure long-term budget balance:
τ
e,t
= (¯r + σ
τ
)(a
g
t
¯a
g
), (20)
When the stock of bonds outstanding exceeds its steady state level, taxes are raised to pay
interest and a portion σ
τ
of the outstanding bonds. Given this fiscal feedback rule for taxes,
government debt then evolves according to (19).
3.5 Equilibrium
We can now define a linearized perfect-foresight transition equilibrium in this economy.
6
Definition 1. Given paths of exogenous shocks {m
t
, η
t
}
t=0
, a linearized perfect foresight equi-
librium is a set of government policies {i
t
, τ
x,t
, τ
e,t
, a
g
t
}
t=0
and a set of aggregates {c
t
, y
t
, a
t
, π
t
,
r
t
, q
b
t
, q
k
t
, w
t
, `
t
}
t=0
such that:
1. The path of aggregate consumption {c
t
}
t=0
is consistent with the linearized aggregate con-
sumption function (18), and the path of household asset holdings {a
t
}
t=1
is consistent with
the budget constraint (8), aggregated across households. a
0
is determined by the existing
portfolios of households now valued at date-0 asset prices.
2. The real wage satisfies (10).
3. The paths {π
t
, y
t
, η
t
}
t=0
are consistent with the Phillips curve (12).
6
All statements in Definition 1 thus refer to the linearized versions of the relevant model equations.
17
4. The paths of {`
t
, y
t
}
t=0
satisfy the aggregate production function (13).
7
5. The asset returns {r
t
, q
b
t
, q
k
t
}
t=0
satisfy (14), (15), and (16).
6. The evolution of government debt a
g
t
and the endogenous component of transfers τ
e,t
are
consistent with the budget constraint (19) and fiscal rule (20).
7. The output and asset markets clear, so y
t
= c
t
+ δk and (a
t+1
a
g
t+1
)/(1 + r
t
) = q
k
t
k.
In this definition, the policy instrument paths {i
t
, τ
x,t
} are taken as given. Sections 4 and 5
will describe optimal policy problems and so discuss how these instruments are determined.
Equilibrium characterization. We can reduce Definition 1 to a small number of linear
relations. Lemma 1 provides this more compact characterization of equilibrium dynamics.
Lemma 1. Given paths of shocks {m
t
, η
t
}
t=0
and government policy instruments {i
t
, τ
x,t
}
t=0
,
paths of aggregate output and inflation {y
t
, π
t
}
t=0
are part of a linearized equilibrium if and
only if
b
π
π
π = κ
b
y
y
y + β
b
π
π
π
+1
+ ψ
b
η
η
η, (21)
b
y
y
y =
˜
C
y
b
y
y
y +
˜
C
π
b
π
π
π +
˜
C
i
b
i
i
i +
˜
C
x
b
τ
τ
τ
x
+ C
m
b
m
m
m, (22)
where the linear maps {
˜
C
y
,
˜
C
π
,
˜
C
i
,
˜
C
x
} are defined in Appendix D.1, and
b
π
π
π
+1
= (π
1
, π
2
, . . . ).
Lemma 1 reduces the complexity of the equilibrium definition in Definition 1 to two
equations: the Phillips curve (21) (which is simply a stacked perfect-foresight version of the
original relation (12)); and the IS curve (22), which differs from the consumption function
(18) chiefly in that it imposes output market-clearing, asset revaluation effects, and feed-
back effects through the government budget to the endogenous component of transfers, τ
e,t
.
Together, these two equations fully characterize the evolution of output and inflation given
exogenous non-policy shocks {m
t
, η
t
}
t=0
and policy choices {i
t
, τ
x,t
}
t=0
.
Discussion. How does our model differ from the canonical representative agent New Key-
nesian models (Gal´ı, 2015; Woodford, 2003)? Positively, the main change is that a simple
7
Note that we drop the efficiency loss term d
t
since it is of second order, and thus does not affect a
first-order approximation of the production function around a zero inflation steady state (see Gal´ı, 2015).
Price dispersion will, however, affect the social welfare function in Section 5.
18
aggregate Euler equation,
by
t
=
1
˜γ
(
b
i
t
bπ
t+1
) + by
t+1
, (23)
is now replaced by the more general aggregate demand block (22).
8
Inequality thus affects
the aggregate dynamics of our economy in response to shocks and policy actions only through
the demand side, with supply—in particular the Phillips curve (21)—kept as in standard
representative-agent models. To arrive at this clean separation, our assumptions on union-
intermediated labor supply (see Section 3.2) are central. We adopt this approach because the
demand-side effects of household heterogeneity are the focus of the recent HANK literature
(e.g., see Kaplan et al., 2018; Auclert et al., 2018).
9
Normatively, household inequality may
affect social welfare functions and thus change policymaker objectives.
The remainder of the paper studies the implications of these two changes for optimal
policy design. First, in Section 4, we isolate the role of changes in propagation by studying a
dual-mandate optimal policy problem. Then, in Section 5, we turn to the Ramsey problem,
thus also allowing the planner objective to change.
4 Optimal dual-mandate policy
We begin our optimal policy analysis by studying the problem of a conventional dual-mandate
policymaker; that is, a policymaker that simply seeks to stabilize fluctuations in inflation and
the output gap. In the context of the structural model of Section 3, such a loss function is ad
hoc, but we nevertheless find it interesting, for two reasons. First, it is conceptually useful—
it allows us to split the effects of inequality on policy design into the role of propagation
(which we study here) and loss function (which we will study in Section 5). Second, it is
arguably policy relevant, as real-world central banks are often mandated to achieve these
types of simple objectives.
We begin in Section 4.1 by stating the optimal policy problem in linear-quadratic form.
We then in Section 4.2 characterize the problem’s solution in the form of a forecast target
criterion. Finally, in Section 4.3, we present some quantitative explorations, leveraging
in particular the connection between our optimal policy “sufficient statistic” formulas and
empirical evidence on the propagation of identified policy shocks.
8
Equation (23) combines the standard consumption Euler equation with EIS 1 with the log-linearized
aggregate resource constraint ˆc
t
= (¯c/¯y)ˆy
t
. The parameter ˜γ that appears in (23) is then γ¯c/¯y.
9
As an added benefit, our union structure with uniform labor rationing allows us to sidestep counterfactual
cross-sectional labor supply responses to policy interventions (see Auclert et al., 2020).
19
4.1 The optimal policy problem
We consider a policymaker with ad hoc objective function
L
DM
X
t=0
β
t
λ
π
bπ
2
t
+ λ
y
by
2
t
. (24)
(24) is a dual-mandate loss function: the policymaker wishes to stabilize inflation and output
around the deterministic steady state, with weights λ
π
and λ
y
, respectively.
For most of this section, we focus on the optimal setting of nominal interest rates i
t
, with
only brief reference to optimal stimulus check policy.
10
The policymaker sets nominal interest
rates to minimize (24) subject to the equilibrium constraints embedded in Definition 1. By
Lemma 1, we can reduce these two constraints to two simple relationships: the Phillips curve
(21) and the IS curve (22). This optimal policy problem is thus a minimal departure from
optimal policy analysis in conventional representative-agent environments: the loss function
(by assumption) and the supply side are unaffected, while the demand constraint changes
from a simple Euler equation as in (23) to the richer demand relation (22).
Note that so far the constraints of this policy problem take the form of our general linear
constraint (2). By the arguments in McKay & Wolf (2023), we can equivalently re-write this
constraint set in impulse response space, thus giving our alternative formulation (3).
11
For
future reference, we write the constraints in impulse response space as
b
π
π
π = Θ
π,i
b
i
i
i + Θ
π,x
b
τ
τ
τ
x
+ Θ
π
b
η
η
η + Θ
π,m
b
m
m
m, (25)
b
y
y
y = Θ
y,i
b
i
i
i + Θ
y,x
b
τ
τ
τ
x
+ Θ
y
b
η
η
η + Θ
y,m
b
m
m
m. (26)
Computational details. Solving the dual-mandate optimal policy problem is straight-
forward. Key to this computational simplicity is that the maps characterizing the linear-
quadratic problem—either the
˜
C’s in the original constraint formulation or the Θ’s in the
equivalent impulse response space formulation—can be obtained straightforwardly as a side-
product of standard sequence-space solution output, following the methods developed in
10
To be more precise, we study the optimal monetary policy problem of setting i
t
, with τ
x,t
= 0 in the
background. There are, of course, still effects of fiscal adjustments through the endogenous fiscal rule (20).
11
In McKay & Wolf (2023), we close the model with a determinacy-inducing policy rule for the policy
instruments, here i and τ
x
. The dynamic causal effect matrices in (25) - (26) are then defined as impulse
response matrices for shocks to the baseline rule. For example, if the inflation and interest rate impulse
response matrices to monetary shocks to the base rule are denoted
˜
Θ
π,i
and
˜
Θ
i,i
, then Θ
π,i
˜
Θ
π,i
˜
Θ
1
i,i
. By
the arguments of McKay & Wolf (2023) this re-writing is without loss of generality.
20
Auclert et al. (2021). It thus follows that optimal policy analysis in the dual-mandate case
comes at essentially zero additional computational cost: if a researcher can solve her HANK
model given a policy rule, then she is only a trivial linear-quadratic problem away from also
obtaining an optimal policy rule for a given quadratic loss.
4.2 Policy rule irrelevance
Our first main result is that, under mild regularity conditions on the linear map
˜
C
i
—i.e., the
mapping from nominal interest rate paths to net excess consumption demand in the gener-
alized “IS” curve (22)—, the optimal monetary policy forecast target criterion is completely
unaffected by household heterogeneity.
Proposition 1. Let
b
c
c
c be a path of household consumption with zero net present value, i.e.,
P
t=0
1
1+¯r
t
bc
t
= 0. If, for any such path
b
c
c
c, we have that
b
c
c
c image(
˜
C
i
), (27)
then the optimal monetary policy rule for a dual-mandate policymaker with loss function (24)
can be written as the forecast target criterion
λ
π
bπ
t
+
λ
y
κ
(by
t
by
t1
) = 0, t = 0, 1, . . . (28)
Recall that our expressions for general linear-quadratic policy problems derived in Sec-
tion 2 immediately yield the optimal dual-mandate policy target criterion as
λ
π
· Θ
0
π,i
·
b
π
π
π + λ
y
· Θ
0
y,i
·
b
y
y
y = 0
0
0. (29)
The proof of Proposition 1 leverages the structure of our model to turn the general expression
(29) into the simple rule (28). Intuitively, while (29) is written in terms of the absolute causal
effects of interest rate policy on output and inflation, (28) is the analogue in terms of relative
causal effects: if monetary policy moves inflation at date t by one unit, then it changes output
at t and t 1 by
1
κ
and β
1
κ
units, respectively. Since the policymaker herself discounts at
rate β, we find that the complicated relation (29) collapses to the simple one (28).
Importantly, the rule (28) is exactly the same as in conventional representative-agent
optimal policy analyses (Gal´ı, 2015; Woodford, 2003). The logic underlying this result is as
follows. In the familiar representative-agent policy problem, any desired path of output and
21
inflation that is consistent with the Phillips curve can be implemented through a suitable
choice of interest rates. The IS curve is thus a slack constraint: the policymaker picks the best
output-inflation pair subject to the NKPC, and then sets interest rates residually to deliver
the required time path of demand. Condition (27) is precisely enough to ensure that this logic
carries through in our environment with household heterogeneity. In words, the condition
says that, through manipulation of short-term nominal interest rates, the policymaker can
engineer any possible net excess demand path with zero net present value. The proof of
Proposition 1 reveals that this is sufficient to ensure that any desired output-inflation pair
consistent with the Phillips curve (21) is in fact implementable. But then, with the Phillips
curve as the supply side of the economy not depending on household inequality, we find that
the target criterion is the same as in conventional representative-agent models.
The implementability condition (27) is discussed further in Wolf (2021). That paper
shows—analytically in simple models, and numerically in heterogeneous-agent environments—
that interest rate policies are indeed generally flexible enough to induce every possible zero
net present value path of aggregate net excess demand. The irrelevance of household hetero-
geneity for optimal forecast criteria is thus a robust feature of HANK-type environments.
Implications for monetary policy practice. The upshot of Proposition 1 is that,
independently of the non-policy shocks hitting the economy, under optimal dual-mandate
policy, the equilibrium paths of output and inflation will be unaffected by household het-
erogeneity and thus equal to those in a standard representative-agent economy. The only
possible effect of heterogeneity is to change the instrument paths—i.e., the current and fu-
ture values of nominal interest rates—required to attain those desired output and inflation
paths. We conclude that the positive implications of household heterogeneity have a rather
limited effect on the practice of conventional dual-mandate policymakers: they can continue
to set their instruments to bring projections of macroeconomic outcomes in line with target,
exactly as done in standard practice of flexible inflation targeting.
12
Optimal stimulus checks. Proposition 1 only considers the first instrument available to
our policymaker: nominal interest rates. Results for stimulus checks follow immediately from
12
Bernanke (2015) succinctly summarizes the salience of this perspective for Federal Reserve policymaking
practice: “The Fed has a rule. The Fed’s rule is that we will go for a 2 percent inflation rate. We will go for
the natural rate of unemployment. We put equal weight on those two things. We will give you information
about our projections about our interest rates. That is a rule and that is a framework that should clarify
exactly what the Fed is doing.”
22
Wolf (2021), who identifies conditions under which interest rate and stimulus check policies
can implement the same sequences of aggregate output and inflation. More formally, it
follows from his results that, if
b
c
c
c image(
˜
C
τ
) (30)
for all sequences
b
c
c
c with zero net present value, then stimulus check policies can also im-
plement the target criterion (28), just like conventional monetary policy. This alternative
implementability condition (30) is again generally satisfied in HANK-type environments.
It follows from the previous discussion that the two policy instruments are perfect sub-
stitutes, and so that the solution to the problem of choosing both instruments to minimize
(24) is indeterminate—multiple paths of the two policy instruments are consistent with the
optimal outcomes for output and inflation. One way to break this indeterminacy is to in-
troduce further constraints on instruments, e.g. a lower bound on nominal interest rates.
Section 5 considers an alternative resolution to this indeterminacy: a richer loss function.
4.3 Quantitative analysis & connection to empirical evidence
We have seen that household heterogeneity does not affect the optimal inflation and output
gap outcomes implemented by a dual-mandate policymaker. Heterogeneity could, however,
in principle affect the time paths of nominal rates required to achieve those optimal outcomes.
This section leverages the close connection between our “sufficient statistics” characterization
of optimal policy and empirical evidence on policy shock propagation to further argue that,
in quantitatively relevant models, the effects of household heterogeneity on optimal rate paths
are likely to be modest in scope.
Exact instrument path irrelevance. We begin with another exact irrelevance result,
building closely on McKay & Wolf (2023). For this result, it will prove convenient to re-state
(5) for the optimal policy instrument path, here specialized to the policy instrument i and
for the policy targets x = (y, π)
0
, in response to some generic set of shocks ε
ε
ε:
b
i
i
i
Θ
0
x,i
W
x,i
1
×
Θ
0
x,i
W
x,ε
· ε
ε
ε
. (31)
Equation (31) states that we can recover the interest rate path required to implement the
optimal policy rule (28) from the absolute causal effects of interest rate policy on output and
inflation. It thus has the following important implication. Any two models—say HANK and
RANK—that agree on (i) the effects of a given shock ε
ε
ε on output and inflation, Θ
y
· ε
ε
ε and
23
Figure 1: Optimal dual-mandate policy response to a cost-push shock. The RANK model replaces
our general “IS” curve (22) with the simple textbook Euler equation (23). We then calibrate all
parameters as in the headline HANK model (see Section 5.3), with one exception: we set the
elasticity of intertemporal substitution (EIS) to generate the same peak response of output to an
identified monetary shock as in the HANK model. In practice, the EIS is little changed.
Θ
π
· ε
ε
ε, and (ii) the effects of interest rate changes on output and inflation, Θ
y,i
and Θ
π,i
,
will necessarily agree on the optimal interest rate path
b
i
i
i
.
We view this irrelevance result as informative because its ingredients are measurable. In
particular, the dynamic causal effects of interest rate changes on aggregate outcomes—that
is, elements of Θ
y,i
and Θ
π,i
—are the estimands of a large empirical literature on monetary
policy shocks (Ramey, 2016). Structural models are often calibrated or estimated to be
consistent with estimates from this literature, which leads them to approximately agree for
the entire matrices Θ
y,i
and Θ
π,i
. The degree to which such quantitative models can disagree
on optimal dual-mandate interest rate paths is thus limited by the empirical evidence.
Quantitative illustration. We close with a quantitative illustration of the analytical
results presented so far. To this end, we study optimal dual-mandate monetary policy in
response to a cost-push shock η
t
in a calibrated version of our HANK model. Details of the
calibration are postponed until Section 5.3; for purposes of the discussion here, it suffices
to note that the model has been parameterized to in particular be consistent with empirical
evidence on the output gap and inflation effects of monetary policy shocks. We then contrast
aggregate outcomes in this economy with those in an analogous RANK model, calibrated
similarly to be consistent with empirical policy shock evidence.
Results are displayed in Figure 1. We emphasize the following two main takeaways, both
consistent with our analytical discussion in the earlier parts of this section. First, as shown
24
in Proposition 1, the output gap and inflation paths are exactly the same. The optimal
trade-off between inflation and output is fully governed by the Phillips curve, which itself
is not affected by household inequality. Second, the nominal interest rate path required to
implement the optimal outcome is quite similar across the two models. Intuitively, since by
construction both models match the same evidence on the transmission from interest rate
changes to output and inflation, the interest rate movements that achieve the policymaker’s
desired output and inflation responses to the shock cannot be too dissimilar.
13
5 Optimal Ramsey policy
We now turn to the optimal policy problem of a Ramsey planner. Unlike our ad hoc dual-
mandate loss function of Section 4, this planner’s objective is directly affected by household
inequality, reflecting a desire to dampen the distributional consequences of aggregate shocks.
Most of our analysis in this section again focuses on optimal monetary policy.
The remainder of the section proceeds in three steps. First, in Section 5.1, we show how
to express the optimal policy problem in our linear-quadratic form. Second, in Section 5.2
we present general analytical results and discuss some instructive analytical special cases.
Finally, in Sections 5.3 and 5.4, we turn to quantitative analysis, leveraging our analytical
“sufficient statistics” formulas to connect to empirical evidence to compare our findings with
those of prior work.
5.1 The optimal policy problem
We consider a conventional Ramsey planner that aggregates utilities of the households pop-
ulating the economy. This section presents a linear-quadratic version of this policy problem.
As is typical in the optimal policy literature (e.g. see the discussion in Giannoni & Wood-
ford, 2002), doing so requires efficiency of the steady state. In the representative agent
context, efficiency requires that the level of production is optimal. Here we also require that
the planner finds the steady-state cross-sectional distribution of consumption desirable. Our
optimal policy analysis will therefore capture an insurance motive against fluctuations in
consumption shares, but it will not seek to change the long-run distribution of consumption.
13
As discussed above, the two interest rate paths would agree exactly if the two models were to agree
on all of Θ
y,i
and Θ
π,i
. Our calibration instead only ensures that peak output and inflation responses in
response to a particular, transitory interest rate movement align. This materially limits the extent of—but
does not fully eliminate—possible disagreement in nominal rate paths.
25
Loss function. To state the loss function, it will prove convenient to describe an individ-
ual’s outcomes in terms of their history of idiosyncratic shocks; that is, we replace c
it
with
ω
t
(ζ
t
i
)c
t
where ζ
t
i
(ζ
it
, ζ
it1
, ζ
it2
, · · · ) is individual i’s history of idiosyncratic shocks and
ω
t
(ζ
t
i
) c
it
/c
t
is their share of aggregate consumption.
14
Letting Γ(ζ) denote the (station-
ary) distribution of such histories (with ζ denoting a generic realization of a history), we can
write the social welfare function as
V
HA
=
X
t=0
β
t
Z
ϕ(ζ)
"
(ω
t
(ζ)c
t
)
1γ
1
1 γ
ν (`
t
)
#
dΓ(ζ), (32)
where ϕ(ζ) is a Pareto weight on the utility of households with history ζ.
In keeping with optimal (monetary) policy analysis in standard representative-agent en-
vironments (Woodford, 2003), our objective is to evaluate the social welfare function (32)
to second order. To this end, a first-order approximation to aggregate equilibrium dynamics
suffices only if the expansion point (i.e., the deterministic steady state) is efficient. Without
household heterogeneity, a simple production subsidy is sufficient to ensure this. With house-
hold heterogeneity, however, we now also require the consumption shares of all households
to be optimal. In principle there are two ways of ensuring this: either the steady-state fiscal
tax-and-transfer system achieves the optimal level of insurance given the planner weights
ϕ(), or the planner weights are set residually so that the implied steady-state distribution
of consumption given a tax-and-transfer system is optimal. We adopt the second approach,
following the inverse optimal taxation literature (e.g. Heathcote & Tsujiyama, 2021). Our
preference for this approach reflects the overarching focus of this paper: we ask how cyclical
policy tools should be manipulated to respond to cyclical changes in inequality, leaving the
long-run steady state outside of the purview of our analysis.
Appendix D.3 presents our assumptions on the production subsidy and policymaker pref-
erence weights that ensure efficiency of the deterministic steady state. Given those assump-
tions, a second-order approximation of (32) around the efficient steady state then gives the
following characterization of the policymaker loss function.
Proposition 2. To second order, the social welfare function V
HA
is proportional to −L
HA
,
given by
L
HA
X
t=0
β
t
bπ
2
t
+
˜κ
¯η
by
2
t
+ λ
ω
Z
bω
t
(ζ)
2
¯ω(ζ)
dΓ(ζ)
, (33)
14
Note that this is without loss of generality, as individuals in our model are ex ante identical, so their
outcomes only differ due to different histories of shocks.
26
where bω
t
(ζ) = ω
t
(ζ) ¯ω(ζ) and ¯ω(ζ) is the steady-state consumption share of an individual
with history ζ and ˜κ and λ
ω
are composite parameters defined in Appendix D.3.
Note that, in the representative-agent analogue of our economy (as discussed in Sec-
tion 3.5), the loss function would feature the same first two terms, as already well-known
from prior work (Woodford, 2003).
15
Our analysis reveals that household heterogeneity adds
a third, inequality-related term, with the planner wishing to stabilize the consumption shares
of everyone in the economy.
How does the inequality term in (33) fit into the linear quadratic framework in Section 2?
Moving to a sequence-space formulation, we can write the loss as
L
HA
= λ
π
b
π
π
π
0
W
b
π
π
π + λ
y
b
y
y
y
0
W
b
y
y
y +
Z
λ
ζ
b
ω
ω
ω(ζ)
0
W
b
ω
ω
ω(ζ)dΓ(ζ), (34)
where λ
π
= 1, λ
y
=
˜κ
¯η
, and λ
ζ
λ
ω
/¯ω(ζ). The consumption share for each idiosyncratic
household history thus emerges as a separate target variable for the policymaker. We will
discuss our approach to computation of (34) later in this section.
16
Constraints. The constraints of the optimal policy problem characterize the evolution
of policymaker targets—π, y, and the consumption shares ω(ζ)—as a function of exogenous
shocks and policy choices. As we discussed in Section 4.1, the evolution of aggregate output
and inflation are governed by the Phillips curve (21) and the IS curve (22). It then remains
to describe the evolution of the inequality term in (33). In Appendix C.2, we establish that,
to first order, we can write the consumption share for a household with specific history ζ as
b
ω
ω
ω(ζ) = Ω
ω(ζ),x
b
x
x
x, ζ (35)
where x
x
x = (y
y
y, r
r
r,τ
τ
τ
x
,τ
e
τ
e
τ
e
,m
m
m, π
0
, q
k
0
, q
b
0
) stacks the prices and shocks that affect the household de-
cision problem, and the maps
ω(ζ),x
give the derivatives of consumption shares with respect
to these aggregate variables. The intuition for (35) is the same as that for the aggregate
consumption function—an individual household’s consumption, given their history of id-
iosyncratic shocks, evolves over time as a function of the aggregate inputs to the household
15
When δ = 0 (so there is no investment spending) the composite term ˜κ reduces to κ, and so the first
two terms in the loss function are exactly the same as the familiar ones from the literature.
16
Note that, technically, (34) does not immediately fit into our framework in Section 2.1 since the objective
here features an integral (rather than a simple sum). Our approach to computation will consider an equivalent
formulation of the problem with finitely many policy targets.
27
consumption-savings problem. By the proof of Lemma 1, we can obtain these inputs as a
function of exogenous shocks, policies, and aggregate output and inflation paths.
For future reference it will be useful to note that, again following McKay & Wolf (2023),
we can alternatively re-write this constraint in impulse response space, solving out the de-
pendence of consumption shares on endogenous aggregates:
b
ω
ω
ω(ζ) = Θ
ω(ζ),i
b
i
i
i + Θ
ω(ζ),x
b
τ
τ
τ
x
+ Θ
ω(ζ)
b
η
η
η + Θ
ω(ζ),m
b
m
m
m. ζ (36)
Summary & computational details. The Ramsey planner chooses paths of the two
available policy instruments—nominal interest rates i
i
i and the exogenous component of trans-
fers τ
τ
τ
x
—to minimize the derived loss function L
HA
subject to the same two constraints as
before, (25) and (26), as well as the evolution of the inequality term, (36).
To computationally evaluate the more complicated loss function (34) and the associated
constraints, we show in Appendix C.2 that the Ramsey loss can be re-written as
L
HA
=
b
π
π
π
0
W
b
π
π
π +
˜κ
¯η
b
y
y
y
0
W
b
y
y
y + λ
ω
b
x
x
x
0
Q
b
x
x
x (37)
where Q is a linear map, defined in Appendix C.2.
17
The alternative formulation in (37)
reflects the simple insight that it is always possible to re-write the loss coming from cross-
sectional inequality as a function of the (small number of) inputs to the consumption-savings
problem—income, interest rates, asset prices, taxes, and shocks. With the re-written loss
function (37), the relevant constraints are then simply the equilibrium dynamics of x
x
x. Com-
putation is thus again straightforward: Appendix C.2 discusses how to recover the map Q,
while the coefficient matrices for all constraints are again immediate from standard sequence-
space solution output, exactly as in the dual-mandate problem considered before. To summa-
rize, relative to solving a HANK model given a policy rule, the only additional computational
work needed to solve a Ramsey optimal policy problem is the one-time computation of the
auxiliary linear map Q.
5.2 Optimal policy rules
Having expressed the optimal Ramsey problem in linear-quadratic form, we can now leverage
the results of Section 2 to provide a general characterization of optimal Ramsey policy rules.
17
We note that the linear map Q in (37) is in general not diagonal. Thus Appendix C.2 also extends the
results from Section 2 to the case with such interaction terms.
28
The optimal Ramsey monetary policy rule is given as
Θ
0
π,i
W
b
π
π
π +
˜κ
¯η
Θ
0
y,i
W
b
y
y
y +
Z
λ
ζ
Θ
0
ω(ζ),i
W
b
ω
ω
ω(ζ)dΓ(ζ) = 0
0
0, (38)
where as before the matrices Θ
,i
collect the dynamic causal effects of interest rate move-
ments on the various policymaker targets—our “sufficient statistics” for optimal policy. In
particular, we see that the first two terms in (38) are identical to the optimal dual-mandate
rule (29), just now with weights derived from policymaker preferences (rather than exoge-
nously assumed). The novel third term—which reflects the planner’s distributional insurance
concerns—collects the causal effects of interest rate movements on consumption shares. Pro-
ceeding analogously for stimulus checks we find the optimal rule
Θ
0
π
x
W
b
π
π
π +
˜κ
¯η
Θ
0
y
x
W
b
y
y
y +
Z
λ
ζ
Θ
0
ω(ζ)
x
W
b
ω
ω
ω(ζ)dΓ(ζ) = 0
0
0, (39)
As discussed in Wolf (2021) and also briefly in Section 4.2, nominal interest rate and stimulus
check policy in our environment can implement exactly the same output-inflation allocations,
so the first two terms in (38) and (39) will reflect identical aggregate stabilization objectives.
The distributional terms Θ
ω(ζ),i
and Θ
ω(ζ)
x
, on the other hand, will generally differ, thereby
opening the door for interest rate and stimulus check policies to be useful complementary
tools for aggregate stabilization. We will return to this observation in Section 5.4. Overall,
(38) and (39) fully characterize joint optimal monetary-fiscal policy.
The key takeaway from the characterizations in (38) and (39) is that optimal policy rules
in the general Ramsey problem deviate from the dual-mandate rules discussed in Section 4
if and only if the policy instruments affect cross-sectional consumption inequality. On the
one hand, if monetary policy is distributionally neutral—as is for example the case in the
environment of Werning (2015) (see Appendix A.2 for details)—then household inequality
does not affect the optimal rule: setting Θ
ω(ζ),i
= 0
0
0 and simplifying (38), we find that the
optimal Ramsey rule continues to take a standard dual-mandate form
18
bπ
t
+
1
¯η
(by
t
by
t1
) = 0, t = 0, 1, . . . (40)
18
Bilbiie (2021) makes an analogous argument in a two-agent context, showing that, if monetary policy
does not redistribute between spenders and savers, then the optimal policy rule is not affected by inequality
concerns. Our analysis reveals that the conditions underlying Werning’s aggregation result are precisely
enough to extend this insight to our heterogeneous-agent setting.
29
In this case, monetary policy optimally focuses on aggregate objectives while fiscal policy
takes sole responsibility for all distributional considerations. On the other hand, if interest
rate cuts are strongly progressive (as is the case in Bhandari et al., 2021; avila & Schaab,
2022), then the concerns about inequality as embedded in the policymaker objective (33)
may materially change optimal monetary policy conduct, with insurance concerns swamping
the usual price and output stabilization motives. This discussion suggests the way forward:
to provide a convincing answer to the question of whether household inequality affects the
optimal Ramsey policy, researchers should consider structural models that agree with empir-
ical evidence on the distributional incidence of policy, as embedded in our sufficient statistics
formulas in (38) - (39). The next subsection discusses how we do so.
5.3 Calibration
We seek to calibrate our structural model to deliver empirically realistic distributional effects
of monetary and fiscal stabilization policy. In particular, as revealed by our formulas in (38)
and (39), what ultimately matters are the effects of policy on consumption inequality. Some
prior empirical work has tried to directly estimate those causal effects. The results so far,
however, are not conclusive: while Holm et al. (2021) find that expansionary monetary policy
has moderately pronounced U-shaped effects on consumption in the wealth cross-section (in
Norway), Coibion et al. (2017) report declines in consumption inequality, and Chang &
Schorfheide (2022) conclude that inequality actually even increases somewhat (both in the
U.S.). In addition to cross-country institutional differences, this lack of consensus is likely
to reflect the lack of high-quality, high-frequency consumption data.
In light of these challenges with direct measurement, we instead proceed as follows.
We calibrate the model to empirical evidence on how monetary policy affects the various
components of household income—for which there is much more consensus as we describe
below. We then rely on the structure of our model’s consumption-savings problem to map
these effects on income into effects on consumption—the sufficient statistics that ultimately
enter our optimal policy formulas (38) and (39). The remainder of this section implements
this approach by proceeding in three steps. First, we review the empirical evidence on how
monetary policy affects income and wealth in the cross-section of households. Second, we
calibrate our model so that it captures those channels of redistribution as well as possible.
And third, we use the model to infer the total effects of monetary policy on the consumption
distribution, bringing together the effects from the various channels.
30
The distributional effects of monetary policy. We begin with a brief review of
the most important transmission channels from monetary policy to household balance sheets.
a) Labor income. There is extensive evidence that those with less education, low past earn-
ings, and racial minorities tend to be more exposed to cyclical fluctuations in labor market
conditions (e.g. Okun, 1973; Hoynes, 2000; Guvenen et al., 2014; Patterson, 2022). Turn-
ing to monetary policy more specifically, Guvenen et al. (2014) study earnings dynamics
in the 1979–1983 recession—a contraction that was arguably caused by a large monetary
intervention. They find that individuals with low previous earnings also suffered the
largest earnings losses. Several recent studies combine identified monetary policy shocks
with European administrative data and arrive at very similar conclusions (Andersen et al.,
2021; Amberg et al., 2021; Holm et al., 2021).
In addition to the distribution of labor income, there is also the question of how strongly
total labor income responds to monetary policy. As is well-known, in textbook sticky-
price, flexible-wage models, expansionary monetary policy sharply increases labor income,
thus giving monetary policy power to redistribute between workers and firm owners.
Empirically, on the other hand, studies have found that expansionary policy leads to a
modest rise in real wages and profits.
19
Our model, due to our assumptions on the division
of monopoly profits between workers and firm owners, implies that total compensation
and returns to capital move one-for-one with output, consistent with the evidence.
b) Asset prices. Expansionary monetary policy raises the value of long-duration assets and
leads to considerable capital gains for the owners of those assets. For example, Bauer &
Swanson (2022) estimate that a 100 basis point reduction in short-term nominal interest
rates increases the S&P 500 stock market index by about five percentage points. As the
distribution of wealth is concentrated, these capital gains are very unequally distributed
across the population. Andersen et al. (2021) leverage administrative household-level data
from Denmark to quantify this distributional gradient. They find that, in response to a
one percentage point decrease in the policy rate, asset values increase by around 70 per
cent of total annual disposable household income at the top of the income distribution,
and around 30 per cent at the median.
20
19
See the summary in Christiano et al. (1999). Relatedly, Cantore et al. (2021) find that expansionary
policy leads to a decrease in the labor share—opposite the response in textbook New Keynesian models.
20
Of course, while long-duration real assets increase in value following an expansionary monetary shock,
so too do the values of long-duration liabilities. In particular, the future consumption plans of households
31
c) Nominal wealth redistribution. Surprise inflation will lower the real value of nominal po-
sitions. Doepke & Schneider (2006) report that wealthy households have about 10-15%
of their net worth in nominal positions, while young middle-class households have consid-
erable nominal liabilities in the form of mortgages. The magnitude of the redistribution
that occurs through this channel of course depends on the strength of the inflation re-
sponse to monetary policy. Many empirical studies point to a relatively flat Phillips curve
(see e.g. the review in Mavroeidis et al., 2014), thus implying relatively small inflation
responses to changes in policy. However, we note that the resulting distributional effects
can be permanent if monetary policy allows permanent changes in the price level.
d) Net interest income. Expansionary policy lowers interest rates, thus leading to lower
debt service payments for debtor households and to lower interest income for creditor
households. This channel benefits middle-class households that own a home with a large
mortgage relative to their net worth (see Cloyne et al., 2020; Wong, 2021).
Model calibration. We calibrate our model to capture the just-summarized channels
of monetary policy propagation to consumption inequality.
Income process. Since many of the distributional effects of monetary policy involve
changes in asset values, it is important that the model generates a concentrated dis-
tribution of wealth, mirroring the data. Since standard incomplete-markets models of
the consumption-savings decision struggle to generate sufficiently concentrated wealth
holdings in the top tail of the distribution, we follow Castaneda et al. (2003), Boar
& Midrigan (2020), and Greenwald et al. (2021) in specifying an income process with
superstar earners, allowing us to generate a realistic concentration of wealth.
We assume that the underlying household income state, ζ
it
, follows a two-component
process, with households being either regular workers or high earners. The function Φ
that maps ζ
it
to labor productivity e
it
is parameterized as
log e
it
= log [(1 α)y
t
] + log (ζ
it
) (1 + m
t
+ χ log y
t
) + log ¯e
t
, (41)
where χ controls the sensitivity of income dispersion to the aggregate business cycle,
with a negative χ implying that low-ζ households are more exposed to the cycle, and
become more costly as real interest rates fall. The theoretical ideal to measure the net exposure to interest
rates, accounting for the difference between asset income and planned consumption (see e.g. the discussion
in Auclert, 2019). Our model will embed these effects on the present value of consumption liabilities.
32
Wealth Income
Data Model Data Model
Top 1% 37 27 17 20
Top 5% 65 66 32 32
Top 10% 76 82 43 44
Top 25% 91 96 64 60
Top 50% 99 101 84 77
Table 1: Shares (%) of wealth and income concentrated in the top x% of the distribution. Data
are from the 2019 Survey of Consumer Finance.
¯e
t
is a normalization constant ensuring that
R
e
it
di = (1 α)y
t
at all dates.
21
For
regular workers, the state log ζ
it
follows an AR(1) process. High-earning households
instead can receive one of two (high) levels of earnings. There are thus six parameters
associated with the high-earnings states: the probability of becoming a high earner
of either kind, the earnings level associated with each state, and the probabilities of
transitioning back to being a regular worker.
22
To estimate the full income process we extend the estimation procedure of Guvenen
et al. (2022). Those authors propose a rich parametric income process that allows for
differential exposure to aggregate business-cycle conditions, and then estimate it using
a variety of moments taken from the Social Security Administration data. This strategy
pins down χ in (41) using data on the differential impact of business-cycle episodes
across the income distribution. We extend their procedure by also targeting moments
of the wealth and income distribution, as reported in Table 1. Further details on the
calibration of the income process and its parameters are presented in Appendix B.1.
Asset structure. In the model, households have positions in capital, short-term bonds,
and long-term bonds. We need to calibrate the total supply of each asset class and then
how these assets are distributed in steady state. The steady state portfolios determine
the distribution of capital gains at t = 0, when the economy is subjected to shocks.
We use data on the balance sheets of U.S. households from the Financial Accounts and
the 2019 SCF to impute the holdings of particular asset classes across the household
21
Regular workers share a common value of χ while high earners have χ = 0.
22
To reduce the number of parameters, we assume that there are no transitions between the high-earning
states and that upon exit the individual receives a random draw from the unconditional distribution of
regular worker ζ’s.
33
Holdings by net worth group
Line Category Total
Top Next Next Bottom
1% 9% 40% 50%
1 Real estate and durables 167 24 48 72 23
2 Equity and mutual funds 191 101 66 23 2
3 Currency, deposits, and similar 60 16 23 19 2
4 Govt. and corp. bonds and similar 29 10 11 7 1
5 Pension assets 131 6 63 58 4
6 Mortgage liabilities 49 2 12 24 11
7 Consumer credit and loans 24 1 2 8 12
8 Net worth excluding pension assets 374 147 135 89 4
9 Capital 419 157 135 101 25
10 Short-term bonds -12 1 7 -3 -16
11 Long-term bonds -33 -11 -8 -9 -5
12 Total 374 147 135 89 4
Table 2: Assets and liabilities in percent of GDP with decomposition by net-worth. Lines 1-8
report the observed asset and liability categories from the Distributional Financial Accounts. Lines
9-12 map these assets and liabilities into their model counterparts.
distribution of net worth. The top panel of Table 2 reports data on asset holdings in
coarse net worth bins, taken from the Distributional Financial Accounts (DFAs).
23
An
important takeaway from Table 2 is that pension assets represent 22% of total house-
hold assets. Of these pension assets, 61% are defined benefit entitlements. Given that
we do not explicitly model pensions we must make a choice about how to incorporate
these assets into our calibration. As households are quite insulated from the returns on
these pension assets, we choose to incorporate the pension funds on to the government
balance sheet. The returns on these pension assets are then gradually paid out to
households through adjustments in taxes.
The next step is to map the observed asset categories into the corresponding asset
classes in the model. We treat corporate and non-corporate equity as levered claims
on capital, so each dollar of equity represents $1.32 dollars of capital and a $0.32 debt
position. This leverage ratio reflects the ratio of debt to assets in the non-financial
business sector. Based on the results in Greenwood et al. (2010) we assume that 62%
23
Our imputation procedure yields similar results to the DFAs, but we use the SCF to create portfolios
that are smooth functions of net worth as opposed to the coarse bins shown in Table 2.
34
of this debt is long-term debt. We set the duration of the long-term bond in our model
to 10 years, in line with results for corporate bonds in Guedes & Opler (1996). Next,
we treat mortgages as a 50-50 split of short-term and long-term debt. Even though
many mortgages are 30-year loans, the duration of these loans is much shorter due
to amortization and prepayment (e.g., Greenwald et al., 2021, report a duration of
approximately 5 years). We treat consumer credit as short-term debt. Overall our
choices lead to a capital-output ratio of 4.19, as shown in Table 2. In total, households
have debts of 12% of GDP in short-term bonds and 33% of GDP in long-term bonds.
24
We set the capital share to α = 0.36 and the depreciation rate to δ = 0.01, allowing
us to match a 17% investment rate (the average observed from 1984 to 2019). These
values imply a steady state annual interest rate of 4.5%. Lastly, we set the borrowing
limit—which restricts net worth and not gross positions—to 0.27 times steady state
average income. This value matches the 5th percentile of net worth in our SCF data.
Fiscal system. The U.S. fiscal tax-and-transfer system is reasonably well-approximated
by a uniform baseline lump-sum transfer coupled with a constant marginal tax rate (see
Kaplan et al., 2018). Data on post-government and pre-government income from the
Congressional Budget Office (2019) imply a steady state transfer of 0.17 times average
income, pinning down the ratio ¯τ
x
/¯y in our model.
25
The speed of fiscal adjustment is
controlled by σ
τ
, which we set equal to σ
b
, implying very gradual fiscal feedback.
26
Other parameters. We set the Frisch elasticity to one, and the elasticity of substitution
between labor varieties to six, based on Basu & Fernald (1997). Finally, we set the
coefficient of relative risk aversion γ and the slope of the Phillips curve κ to match
the peak responses of output and inflation to an empirically identified monetary policy
shock, ensuring that our model is not just consistent with distributional incidence, but
also with overall level responses. Again, the details for this internal calibration are
provided in Appendix B.1.
24
As households (excluding pension funds) collectively have negative credit positions, the pension funds
will lend to the households and to the government. As we consolidate the pension funds and government
in our model, it is thus as if the government is lending to the households. See Appendix B.2 for further
description of how we use the SCF to impute asset positions as smooth functions of household net worth.
25
We exclude Medicaid and CHIP benefits from the after-tax income. We then regress after-tax incomes
on before-tax incomes for the first four quintiles of the income distribution. The intercept of this regression
gives us the steady state transfer.
26
We have found that the results are robust to the choice of σ
τ
as long as it takes a low value. Empirically,
changes in public debt are highly persistent, thus implying a low value of σ
τ
.
35
Parameter Description Value Calibration target
Households
{ζ
it
} Income risk process See text
χ Income exposure -2.0 Het. earnings cyclicality
γ Relative risk aversion 1.2 Monetary shock effects
φ Frisch elasticity 1 Standard
β Discount factor 0.984 Asset market clearing
Firms
κ Phillips curve slope 0.022 Monetary shock effects
¯η Labor Substitutability 6 Basu & Fernald (1997)
Asset Structure
α Capital share 0.36 Standard
δ Depreciation rate 0.01 Investment rate
a/¯y Borrowing limit 0.27 Fifth percentile of net worth
σ
b
Long-term bond duration 0.025 Guedes & Opler (1996)
Government
τ
y
Labor tax rate 0.23 Steady-state budget balance
¯τ
x
¯y
Transfer share 0.17 After-tax vs. before-tax income
σ
τ
Tax-debt responsiveness = σ
b
See text
Table 3: Calibration of our quantitative HANK model. The model period is one quarter.
We summarize all parameter values in Table 3.
Consumption responses. Having calibrated our model to be consistent with the most
salient transmission channels of monetary policy to households, we now use the structural
model to map these income responses into the sufficient statistics that ultimately enter our
optimal policy rules—the effects of monetary policy on consumption inequality.
Figure 2 shows the model-implied impact change in consumption across the wealth distri-
bution following a transitory, expansionary monetary policy shock. Since histories ζ map into
differences in income and wealth, we note that this figure is directly informative about the
Θ
ω(ζ),i
’s that enter our optimal policy rules. Our main finding is that households across the
wealth distribution increase their consumption relatively uniformly, implying small effects
on consumption shares—that is, causal effects Θ
ω(ζ),i
that are quite close to zero.
36
Figure 2: Initial response of consumption to an expansionary monetary policy shock across the
distribution of wealth. We simulate the shock we estimate empirically in Appendix B.1.
What explains these rather evenly distributed effects of policy stimulus? On the one end
of the spectrum, low-wealth households tend to also have rather low incomes, their labor
earnings are more exposed to aggregate income, and they are often borrowing-constrained
and so tend to have high marginal propensities to consume (MPCs). As a result, changes
in labor earnings pass through strongly into consumption. Households with a moderate
amount of wealth instead tend to have substantial real estate and mortgage positions (see
the Next 40% column in Table 2), and also relatively moderate MPCs. Finally, high-wealth
households benefit from increases in asset values. While these capital gains are very large, the
pass-through to consumption is relatively weak, given the rather low model-implied MPCs
at the top end of the wealth distribution.
27
Putting all of the pieces together, we arrive at
the relatively uniform effects shown in Figure 2.
27
Note that the earnings of the lowest-wealth group increase by 0.5% of their steady-state income, while
the capital gains of the highest-wealth group amount to more than 20% of steady-state income. The model
is thus fully consistent with the empirical results of Andersen et al. (2021) and Bartscher et al. (2021), who
emphasize that capital gains effects are large relative to labor income effects. While the implications for
short-run consumption are more equal, high-wealth consumption does remain elevated for longer. These
dynamic effects are not visible in Figure 2, but they are incorporated into our analysis of optimal policy.
37
5.4 Applications
We now use the calibrated model as a laboratory to explore optimal Ramsey policy in two
scenarios. As our first application, we consider an aggregate shock with strong distributional
consequences, depressing consumption of the poor relative to that of the rich, akin to the
Covid-19 recession. Second, we return to the inflationary cost-push shock from Section 4.3
a shock that has received much attention in the classic RANK literature as well as the recent
HANK literature (e.g., Bhandari et al., 2021; Acharya et al., 2022). We will use this second
application and the comparison with prior work to in particular highlight the usefulness of
our sufficient statistics approach to optimal policy rule characterization.
Distributional shock. Our first shock is an innovation to m
t
, the exogenous driver of
income dispersion in (41). This shock redistributes labor income from low-income to high-
income households, thus depressing aggregate demand: precautionary savings increase due
to the increase in risk, and so spending falls as income is redistributed towards lower-MPC
households. We study three optimal policy responses to this demand shock: monetary policy
for a dual mandate policymaker; monetary policy for a Ramsey planner; and finally joint
monetary-fiscal policy for a Ramsey planner. Results are reported in Figure 3.
We begin with the optimal dual mandate policy response, displayed as the grey lines
in Figure 3. As aggregate demand falls, the dual-mandate central banker cuts nominal
interest rates to perfectly stabilize output and inflation (the classic “divine coincidence”).
The bottom panel shows how the consumption distribution changes on impact of the shock:
the shock itself redistributes from low-income to high-income households, and monetary
policy—precisely because of how we calibrated the model—does very little to offset that.
We then consider the optimal Ramsey monetary policy that also considers distributional
objectives, depicted as the orange dashed lines in Figure 3. Our headline finding here is that
the optimal policy response is very similar to the dual-mandate policy, with nominal interest
rates now cut somewhat more on impact of the shock but then very closely following the
dual mandate policy. As a result, aggregate output and inflation continue to be stabilized
almost perfectly, and the response of the consumption distribution hardly differs from the
dual-mandate outcome. The intuition is as follows. The original shock m
t
has a strong
distributional tilt, with low-income households losing the most. The Ramsey planner would
like to lean against this redistribution and stabilize consumption shares. Monetary policy,
however, is a rather blunt tool: if the planner were to cut rates by enough to stabilize
consumption at the bottom of the distribution, then consumption of richer households and
38
Figure 3: Optimal policy response to an income distribution shock m
t
. The figure shows the
results from three policy rules: optimal monetary policy for a dual-mandate policymaker (Dual
Mandate), optimal monetary policy for a Ramsey planner that is constrained not to use lump-sum
transfers (Monetary Only), and optimal policy for a Ramsey planner that can use both tools (Joint
Monetary-Fiscal). Transfers are expressed in units of dollars, using the conversion that steady state
per capita GDP is $60,000.
39
so aggregate output and inflation would overshoot significantly—an immediate implication
of the flat distributional effects of monetary interventions shown in Figure 2. We provide a
visual illustration of this point in Appendix Figure B.4: increasing the size of the interest
rate cut to stabilize the consumption of households with low income and low wealth comes
at the cost of over-stimulating the consumption of other households (and thus aggregate
output and inflation).
Finally we turn to the optimal joint monetary-fiscal Ramsey policy, with a planner using
both interest rate policy and stimulus checks to pursue her aggregate and distributional
objectives. Results are displayed as the dark blue lines in Figure 3. Compared to monetary
policy, fiscal stimulus payments are much more progressive, with consumption of low-income
and low-wealth households responding significantly more than that of high-income, high-
wealth households (see Wolf (2021) or Appendix Figure B.3). Intuitively, this reflects both
differences in MPCs as well as differences in income levels, with any given dollar amount of
stimulus checks amounting to a much larger fraction of income at the bottom of the income
distribution. As a result, fiscal stimulus payments are particularly well-suited as a tool to
address shocks that differently affect low-income and high-income households. The results
depicted in Figure 3 are consistent with this intuition: in response to the distributional
shock m
t
, the Ramsey planner can use checks to almost perfectly stabilize aggregate output,
inflation, and consumption inequality. Since stimulus payments both compress inequality
and stimulate aggregate demand, there is little need for the monetary authority to intervene,
with nominal interest rates now responding very little.
Overall, our analysis of optimal Ramsey policy responses to the distributional demand
shock m
t
suggests two broad lessons for optimal cyclical policy design. First, recall from our
prior analysis that a monetary policymaker has an incentive to deviate from the prescrip-
tions of dual mandate policy only if monetary policy can help her to offset the distributional
incidence of the underlying business-cycle shock. With empirical evidence suggesting a rel-
atively uniform distributional incidence of monetary interventions, it follows that monetary
policy is in general rather ill-suited to attain distributional objectives. In particular, sta-
bilizing consumption at the bottom of the distribution may require significant departures
from dual-mandate objectives; if the Ramsey planner is unwilling to accept these departures
(as was the case in our example), then optimal policy will tend to stay close to the dual-
mandate benchmark, as seen here. Second, fiscal stimulus payments and nominal interest
rate policy can be highly complementary policy tools. They are likely to have very different
cross-sectional incidence profiles (see Figure 2 and Figure B.3), and so the optimal mix of
40
Figure 4: Optimal policy response to an inflationary cost-push shock. The figure shows optimal
monetary policy for a planner whose objective is to stabilize output and inflation (Dual Mandate)
and for a Ramsey planner (Monetary Ramsey).
monetary and fiscal policy as a stabilization tool will depend on the distributional incidence
of the shocks hitting the economy.
Cost-push shock. Our second shock is a classic cost-push shock η
t
, as already considered
in our dual-mandate application in Section 4.3. In our model, this cost-push shock introduces
a wedge in the aggregate Phillips curve, thus giving an output-inflation trade-off. Under our
assumptions on the incidence of firm profits, however, this shock does not by itself have
strong distributional implications. For our analysis we restrict attention to monetary policy
alone, comparing only the optimal dual-mandate and Ramsey monetary policy responses.
Results are displayed in Figure 4. We find that the Ramsey policy is nearly identical to the
41
dual-mandate policy. Intuitively, neither the shock nor the monetary policy response has
strong distributional effects, and so there is little reason to deviate from the dual mandate
policy. Furthermore, by the logic discussed above, this result is unlikely to change materially
even if the cost-push did have meaningful distributional effects, simply because monetary
policy would still be ill-suited to offset those.
Our findings contrast with recent arguments made in Bhandari et al. (2021), who con-
clude that distributional considerations may lead to substantial changes in optimal policy
conduct—changing the signs of impulse response functions and the order of magnitude of
inflation volatility. Our analysis differs from theirs both methodologically and due to the as-
sumptions on the economic environment. To demonstrate that the difference in conclusions
is driven to a considerable extent by the economic assumptions, Appendix B.4 repeats our
analysis of the cost-push shock in an alternative model environment that is more similar to
that in Bhandari et al.. In particular, for this alternative environment, we assume that (i) all
monopoly profits are distributed to wealthy firm owners and (ii) there are no other types of
long-duration assets (i.e., the model omits capital and housing-related assets and liabilities).
This change in the model environment has two implications. First, cost-push shocks now
have strong distributional effects: they increase markups, decrease real wages, and increase
the profits flowing to a concentrated group of firm owners. Second, expansionary monetary
policy is also redistributive, but in the other direction—it reduces markups, raises real wages,
and reduces profits. In light of our previous discussion it is thus unsurprising that we find an
optimal Ramsey monetary policy closely echoing the conclusions of Bhandari et al.: while
a dual mandate policymaker hikes interest rates following a cost-push shock, the Ramsey
planner instead sharply lowers them to reduce cross-sectional inequality.
6 Conclusion
Should household inequality affect the conduct of cyclical stabilization policy? The analysis
in this paper suggests the following three main takeaways.
First, for central banks that target standard macroeconomic aggregates (e.g., a traditional
“dual mandate”), household inequality—at least as typically modeled in the burgeoning
HANK literature—is likely to only have rather moderate effects. Analytically, we have
given conditions under which the optimal forecast target criterion of a dual mandate central
banker is unaffected by household inequality, implying that output and inflation outcomes
in response to any shock will be exactly as in analogous representative-agent environments.
42
Empirically, a long literature already estimates the causal effects of interest rate changes
on output and inflation. Models that are consistent with this evidence will yield similar
predictions for the paths of nominal rates necessary to implement a given output and inflation
target—irrespective of whether the model features household inequality or not.
Second, the extent to which distributional objectives shape optimal Ramsey monetary
policy depends crucially on the causal effects of interest rate changes on consumption inequal-
ity. According to our analysis, monetary policy has relatively even effects on consumption
across households. Interest rate policy is thus not a particularly sharp tool to deal with
shocks that disproportionately affect the poor, at least not without substantial costs in
terms of aggregate stabilization.
Third, fiscal stimulus checks—an alternative tool of stabilization policy, used increasingly
frequently in recent decades—promise to be more useful for distributional purposes. Such
stimulus checks achieve aggregate stabilization through insurance at the bottom, thus making
them well suited to address cyclical fluctuations that mostly affect poor households.
43
References
Acharya, S., Challe, E., & Dogra, K. (2022). Optimal monetary policy according to HANK.
Forthcoming American Economic Review.
Alves, F., Kaplan, G., Moll, B., & Violante, G. L. (2020). A further look at the propagation of
monetary policy shocks in hank. Journal of Money, Credit and Banking, 52 (S2), 521–559.
Amberg, N., Jansson, T., Klein, M., & Rogantini Picco, A. (2021). Five facts about the
distributional income effects of monetary policy. CESifo Working Paper.
Andersen, A. L., Johannesen, N., Jørgensen, M., & Peydr´o, J.-L. (2021). Monetary policy
and inequality. Working Paper.
Auclert, A. (2019). Monetary policy and the redistribution channel. American Economic
Review, 109 (6), 2333–67.
Auclert, A., Bard´oczy, B., & Rognlie, M. (2020). Mpcs, mpes and multipliers: A trilemma
for new keynesian models. Technical report, National Bureau of Economic Research.
Auclert, A., Bard´oczy, B., Rognlie, M., & Straub, L. (2021). Using the sequence-space
jacobian to solve and estimate heterogeneous-agent models. Econometrica, 89 (5), 2375–
2408.
Auclert, A. & Rognlie, M. (2018). Inequality and aggregate demand. Technical report,
National Bureau of Economic Research.
Auclert, A., Rognlie, M., Souchier, M., & Straub, L. (2021). Exchange rates and monetary
policy with heterogeneous agents: Sizing up the real income channel. Technical report,
National Bureau of Economic Research.
Auclert, A., Rognlie, M., & Straub, L. (2018). The intertemporal keynesian cross. Technical
report, National Bureau of Economic Research.
Bartscher, A. K., Kuhn, M., Schularick, M., & Wachtel, P. (2021). Monetary policy and
racial inequality. CEPR Discussion Paper No. DP15734.
Basu, S. & Fernald, J. G. (1997). Returns to scale in us production: Estimates and implica-
tions. Journal of political economy, 105 (2), 249–283.
44
Bauer, M. D. & Swanson, E. T. (2022). A reassessment of monetary policy surprises and
high-frequency identification. National Bureau of Economic Research Working Paper.
Benigno, P. & Woodford, M. (2012). Linear-quadratic approximation of optimal policy
problems. Journal of Economic Theory, 147 (1), 1–42.
Bernanke, B. (2015). Remarks at Brookings Institution: Objections to Federal Reserve
Accountability Bill.
Bhandari, A., Evans, D., Golosov, M., & Sargent, T. J. (2021). Inequality, business cycles,
and monetary-fiscal policy. Econometrica, 89 (6), 2559–2599.
Bilbiie, F. O. (2021). Monetary policy and heterogeneity: An analytical framework. Working
Paper.
Bilbiie, F. O., Monacelli, T., & Perotti, R. (2021). Stabilization vs. redistribution: the
optimal monetary-fiscal mix. CEPR Discussion Paper No. DP15199.
Boar, C. & Midrigan, V. (2020). Efficient redistribution. National Bureau of Economic
Research Working Paper.
Boppart, T., Krusell, P., & Mitman, K. (2018). Exploiting MIT shocks in heterogeneous-
agent economies: the impulse response as a numerical derivative. Journal of Economic
Dynamics and Control, 89, 68–92.
Cantore, C., Ferroni, F., & Le´on-Ledesma, M. (2021). The missing link: monetary policy
and the labor share. Journal of the European Economic Association, 19 (3), 1592–1620.
Castaneda, A., Diaz-Gimenez, J., & Rios-Rull, J.-V. (2003). Accounting for the us earnings
and wealth inequality. Journal of Political Economy, 111 (4), 818–857.
Chang, M. & Schorfheide, F. (2022). On the effects of monetary policy shocks on earnings
and consumption heterogeneity. CEPR Discussion Paper No. DP17049.
Christiano, L. J., Eichenbaum, M., & Evans, C. L. (1999). Monetary policy shocks: What
have we learned and to what end? Handbook of Macroeconomics, 1, 65–148.
Cloyne, J., Ferreira, C., & Surico, P. (2020). Monetary policy when households have debt:
new evidence on the transmission mechanism. The Review of Economic Studies, 87 (1),
102–129.
45
Coibion, O., Gorodnichenko, Y., Kueng, L., & Silvia, J. (2017). Innocent bystanders? mon-
etary policy and inequality. Journal of Monetary Economics, 88, 70–89.
Congressional Budget Office (2019). Projected changes in the distribution of household
income, 2016 to 2021. Technical report, Congressional Budget Office. Report.
avila, E. & Schaab, A. (2022). Optimal monetary policy with heterogeneous agents: A
timeless ramsey approach. Working Paper.
De Groot, O., Mazelis, F., Motto, R., & Ristiniemi, A. (2021). A toolkit for computing
constrained optimal policy projections (copps). ECB Working Paper.
Doepke, M. & Schneider, M. (2006). Inflation and the redistribution of nominal wealth.
Journal of Political Economy, 114 (6), 1069–1097.
Farhi, E. & Werning, I. (2019). Monetary policy, bounded rationality, and incomplete mar-
kets. American Economic Review, 109 (11), 3887–3928.
Fern´andez-Villaverde, J., Rubio-Ram´ırez, J. F., & Schorfheide, F. (2016). Solution and
estimation methods for DSGE models. In Handbook of Macroeconomics, volume 2 (pp.
527–724). Elsevier.
Gal´ı, J. (2015). Monetary policy, inflation, and the business cycle: an introduction to the
new Keynesian framework and its applications. Princeton University Press.
Gertler, M. & Karadi, P. (2015). Monetary policy surprises, credit costs, and economic
activity. American Economic Journal: Macroeconomics, 7 (1), 44–76.
Giannoni, M. P. & Woodford, M. (2002). Optimal interest-rate rules. NBER Working Paper.
Greenwald, D. L., Leombroni, M., Lustig, H., & Van Nieuwerburgh, S. (2021). Financial
and total wealth inequality with declining interest rates. National Bureau of Economic
Research Working Paper.
Greenwood, R., Hanson, S., & Stein, J. C. (2010). A gap-filling theory of corporate debt
maturity choice. The Journal of Finance, 65 (3), 993–1028.
Guedes, J. & Opler, T. (1996). The determinants of the maturity of corporate debt issues.
the Journal of Finance, 51 (5), 1809–1833.
46
Guvenen, F., McKay, A., & Ryan, C. (2022). A tractable income process for business cycle
analysis. Working Paper.
Guvenen, F., Ozkan, S., & Song, J. (2014). The nature of countercyclical income risk.
Journal of Political Economy, 122 (3), 621–660.
Heathcote, J. & Tsujiyama, H. (2021). Optimal income taxation: Mirrlees meets ramsey.
Journal of Political Economy, 129 (11), 3141–3184.
Hebden, J. & Winker, F. (2021). Impulse-Based Computation of Policy Counterfactuals.
Working Paper.
Holm, M. B., Paul, P., & Tischbirek, A. (2021). The transmission of monetary policy under
the microscope. Journal of Political Economy, 129 (10), 2861–2904.
Hoynes, H. (2000). The employment, earnings, and income of less skilled workers over the
business cycle. In B. Rebecca & D. Card (Eds.), Finding Jobs: Work and Welfare Reform
(pp. 23–71). Russell Sage Foundation.
Kaplan, G., Moll, B., & Violante, G. L. (2018). Monetary policy according to hank. American
Economic Review, 108 (3), 697–743.
Le Grand, F., Martin-Baillon, A., & Ragot, X. (2022). Should monetary policy care about
redistribution? optimal fiscal and monetary policy with heterogeneous agents. Technical
report, Working Paper, SciencesPo.
Mavroeidis, S., Plagborg-Møller, M., & Stock, J. H. (2014). Empirical evidence on inflation
expectations in the new keynesian phillips curve. Journal of Economic Literature, 52 (1),
124–88.
McKay, A. & Wolf, C. K. (2023). What can time-series regressions tell us about policy
counterfactuals? MIT Working Paper.
Okun, A. M. (1973). Upward mobility in a high-pressure economy. Brookings Papers on
Economic Activity, 4 (1), 207–262.
Patterson, C. (2022). The matching multiplier and the amplification of recessions. Working
Paper.
47
Plagborg-Møller, M. & Wolf, C. K. (2021). Local projections and vars estimate the same
impulse responses. Econometrica, 89 (2), 955–980.
Ramey, V. A. (2016). Macroeconomic shocks and their propagation. Handbook of macroe-
conomics, 2, 71–162.
Svensson, L. E. (1997). Inflation forecast targeting: Implementing and monitoring inflation
targets. European economic review, 41 (6), 1111–1146.
Svensson, L. E. (2005). Monetary policy with judgment: Forecast targeting.
Werning, I. (2015). Incomplete markets and aggregate demand. Working Paper 21448,
National Bureau of Economic Research.
Wolf, C. K. (2021). Interest rate cuts vs. stimulus payments: An equivalence result. National
Bureau of Economic Research Working Paper 29193.
Wong, A. (2021). Refinancing and the transmission of monetary policy to consumption.
Working paper.
Woodford, M. (2003). Interest and prices: Foundations of a theory of monetary policy.
princeton university press.
48
A Supplementary model details
This appendix provides some supplementary details for our structural model. Appendix A.1
begins by further discussing the price-setting problem and deriving the log-linearized Phillips
curve (12), and Appendix A.2 discusses the model’s special case in which monetary policy
is distributionally neutral, following Werning (2015).
A.1 Phillips Curve
Cost minimization by the final goods firm yields the usual demand curve for intermediate
varieties
y
jt
=
p
jt
p
t
η
y
t
. (A.1)
and the usual price index
p
t
=
Z
p
1η
jt
dj
1
1η
. (A.2)
We now turn to the price-setting problem of the intermediate goods firm. We assume
that intermediate goods production is subsidized at gross rate
¯ηΞ
(¯η1)(1τ
y
)
, where ¯η is the
steady state elasticity of substitution between varieties of labor and the term Ξ accounts for
the fact that the social planner may attach a higher value to resources than the labor union
(we derive the precise value of Ξ in Appendix D.3); for the purposes of our analysis here, it
suffices to note that the labor subsidy takes this general form and that it is financed with
a lump-sum tax on the monopolostically competitive firms. The intermediate producer’s
problem is to choose the reset price p
t
and the input mix {k
jt+s
, `
jt+s
}
s0
to maximize
X
s0
β
s
θ
s
¯ηΞ
(¯η 1)(1 τ
y
)
p
t
p
t+s
y
j,t+s
r
k
t+s
k
jt+s
w
t+s
`
jt+s
subject to (A.1), (11) and taking r
k
t+s
and w
t+s
as given. The Lagrangian is
L =
X
s0
β
s
θ
s
"
¯ηΞ
(¯η 1)(1 τ
y
)
p
t
p
t+s
p
t
p
t+s
η
t+s
y
t+s
r
k
t+s
k
jt+s
w
t+s
`
jt+s
49
µ
t+s
p
t
p
t+s
η
t+s
y
t+s
Ak
α
jt+s
`
1α
jt+s
!#
and the first-order conditions are
X
s0
β
s
θ
s
¯ηΞ(η
t+s
1)
(¯η 1)(1 τ
y
)
p
t
p
t+s
η
t+s
y
t+s
p
t+s
=
X
s0
β
s
θ
s
η
t+s
µ
t+s
p
t
p
t+s
η
t+s
1
y
t+s
p
t+s
!
(A.3)
r
k
t+s
= µ
t+s
α
y
jt+s
k
jt+s
(A.4)
w
t+s
= µ
t+s
(1 α)
y
jt+s
`
jt+s
. (A.5)
Taking the ratio of (A.4) and (A.5) we obtain
r
k
t+s
w
t+s
=
α
1 α
`
jt+s
k
jt+s
,
which implies that all firms use the same capital-labor ratio, which by market clearing must
be the aggregate capital-labor ratio. Equations (A.4) and (A.5) imply that capital receives
a share µ
t
α of output and labor receives a share µ
t
(1 α), which leaves a share (1 µ
t
) for
monopoly profits. As we assume workers receive a share 1 α of the monopoly profits, the
total compensation to workers is a share 1 α of output.
Using (11) in (A.5) we obtain
w
t+s
= µ
t+s
(1 α)A
k
jt+s
`
jt+s
α
= µ
t+s
(1 α)A
k
`
t+s
α
,
where the second equality replaces the firm-specific capital-labor ratio with the aggregate
capital-labor ratio by the argument above. Now use (10) to substitute for w
t+s
and log-
linearize to obtain
ˆµ
t+s
= (φ + α)
ˆ
`
t+s
+ γˆc
t+s
,
where φ = v
00
(`)`/v
0
(`) is the inverse Frisch elasticity. To a first-order approximation, the
aggregate production function gives
ˆy
t+s
= (1 α)
ˆ
`
t+s
50
and the aggregate resource constraint gives
ˆy
t+s
=
¯c
¯y
ˆc
t+s
.
Combining the last three equations gives
ˆµ
t+s
=
φ + α
1 α
+
¯yγ
¯c
ˆy
t+s
. (A.6)
Log-linearizing the (A.3) around a zero-inflation steady state yields
bp
t
bp
t
= (1 βθ)
X
s0
β
s
θ
s
ˆµ
t+s
1
¯η 1
bη
t+s
+ bp
t+s
bp
t
Next, we from the definition of the price index have
1 + π
t
p
t
p
t1
=
θ
1
1 θ
θ
p
t
p
t
1η
t
!
1
η
t
1
. (A.7)
Log-linearizing around a zero inflation steady state this gives
bπ
t
= bp
t
bp
t1
=
1 θ
θ
(bp
t
bp
t
) .
Eliminating bp
t
bp
t
, using (A.6), and simplifying, we get
bπ
t
= κby
t
+ ψbη
t
+ βbπ
t+1
where κ =
(1θ)(1βθ)
θ
φ+α
1α
+
¯yγ
¯c
and ψ =
(1θ)(1βθ)
θ(¯η1)
.
A.2 Werning (2015) special case
This section elaborates on our discussion in Section 5.2 of optimal policy in a special case of
our model where monetary policy is distributionally neutral, following Werning (2015). We
proceed in three steps. First, we present the assumptions required to arrive at this special
case. Second, we derive the distributional irrelevance result. And third, we formally state
implications for optimal policy design. Proofs are deferred to Appendices D.4 to D.7.
51
Model assumptions. We consider a special case of our environment, adapted to be to
be consistent with the assumptions in Werning (2015).
First, we further restrict the household consumption-savings problem.
Assumption A.1. Household utility is logarithmic (γ = 1). The distribution of household
earnings e
it
is acylical (i.e., the incidence function Φ from idiosyncratic events to earnings
can be written as
˜
Φ(ζ
it
, m
t
) × (1 α)y
t
so earnings scale with y
t
) and households can self-
insure only through saving, not borrowing (a = 0).
Second, we assume that the asset supply has particular properties.
Assumption A.2. There is no depreciation (δ = 0), there is no net supply of bonds at any
date, and at date-0 capital is the only asset held by households.
The third and final assumption restricts the exogenous component of transfers, τ
x
.
Assumption A.3. In steady-state, the exogenous and endogenous components of govern-
ment transfers to households are zero, i.e. ¯τ
x
= 0 and ¯τ
e
= 0.
Under similar assumptions, Werning (2015) proves that the demand side of the economy
responds to monetary policy exactly like the conventional Euler equation (23). Household
heterogeneity affects the split of the consumption response into indirect income and direct
interest rate effects, but leaves the overall sum unchanged. We will see that the same logic
allows a sharp characterization of optimal policy rules for the Ramsey planner.
The distributional effects of monetary policy. Following steps similar to those
in Werning (2015), we can prove the following useful building block result.
Proposition A.1. Under Assumptions A.1 to A.3, we have that
Θ
ω(ζ),i
= 0
0
0 ζ. (A.8)
In words, changes in nominal interest rates have no effect on the consumption distribution,
at any horizon.
Optimal policy characterization. Combining Proposition A.1 and our general char-
acterizations of optimal forecast targeting policy rules in Section 2.1, it follows immediately
that the presence of the inequality term in the policymaker loss function does not at all affect
the target criterion for optimal monetary policy. Corollary A.1 summarizes this conclusion.
52
Corollary A.1. Under Assumptions A.1 to A.3 and the conditions of Proposition 1, the
optimal monetary policy rule for a Ramsey policymaker with loss function (33) can be written
as the forecast target criterion
bπ
t
+
1
¯η
(by
t
by
t1
) = 0, t = 0, 1, . . . (A.9)
Corollary A.1 formalizes our intuitive discussion in Section 5.2. Finally we also note the
following implication of these results: since monetary policy stabilizes output and inflation as
well as possible, and since fiscal policy has no additional scope to help with this stabilization
(recall Section 4), it follows that fiscal transfer policy is exclusively concerned with inequality
stabilization, with the optimal transfer target criterion given as
Z
λ
ω(ζ)
Θ
0
ω(ζ)
x
W
b
ω
ω
ω(ζ)dΓ(ζ) = 0
0
0. (A.10)
Corollary A.2 summarizes this conclusion.
Corollary A.2. Under the optimal joint fiscal-monetary policy, transfers are set following
the target criterion (A.10), minimizing the inequality term in the loss function (33). Mon-
etary policy is set residually to enforce the target criterion (A.9), thus attaining the same
paths of aggregate inflation and output as under optimal monetary policy alone.
We thus achieve a strict separation of monetary and fiscal instruments. Corollary A.2
is related to the results in Bilbiie et al. (2021): while those authors find that transfers can
be set to perfectly stabilize inequality between two groups of households, transfers in our
HANK model are set to stabilize the general inequality term as well as possible. In both
cases, given fiscal stabilization of inequality, conventional monetary policy then implements
the aggregate allocations familiar from standard representative-agent analysis—in their case
because inequality is already perfectly stabilized; in our case because conventional monetary
policy cannot help any further with the inequality-related loss.
53
B Supplementary material for Sections 5.3 and 5.4
This appendix presents supplementary material for our quantitative optimal Ramsey policy
analysis. Appendix B.1 begins with further details on model calibration, while Appendix B.3
provides various additional results for our particular business-cycle shock experiments.
B.1 Calibration
Income process. We use a similar estimation procedure as Guvenen et al. (2022) in
fitting the parameters of the income process. We briefly summarize the moments that are
targeted here and describe how we deviate from Guvenen et al. while referring the reader
to that paper for most of the details.
To concisely describe the income process, it is useful to adopt the notation that ζ
it
Z
R
for regular workers and ζ
it
{z
H
1
, z
H
2
} for high-earners. We estimate the following income
process:
log e
it
= log [(1 α)y
t
] + log (ζ
it
)
1 + m
t
+ χI
[ζ
it
∈Z
R
]
log y
t
+ log ¯e
t
+ µ
i
+ ξ
T
it
As described by Guvenen et al., over the lifecycle, the cross-sectional variance of earnings
grows almost linearly in the age of a cohort. In the absence of dispersion in the deterministic
component of individual income profiles, this implies near-random walk behavior in individ-
ual earnings. To have a stationary distribution of income, we fix ρ
ζ
at a value less than
one, setting ρ
ζ
= 0.9
1/4
.
28
ξ
T
it
is a transitory income shock that follows a two-state Markov
chain. We fix the parameters of this Markov chain to mimic the “non-employment” shock
in Guvenen et al..
29
Finally µ
i
is an individual fixed effect.
If ζ
it1
Z
R
, with probability Π
RR
the worker remains a regular worker and ζ
it
follows
log ζ
it
= ρ
ζ
log ζ
it1
+ ξ
it
ξ
it
N(0, σ
2
ξ
).
In principal, the Gaussian shocks imply the support for log ζ
it
for regular workers is all
28
Allowing for heterogeneous income profiles, one finds a lower value of ρ
ζ
.
29
We include this shock in our income process estimation so that the persistent shocks ξ
it
are not forced
to account for all features of the income data. For the sake of parsimony, however, we do not include this
shock in quantitative model analysis. We have experimented with including this shock and found the results
to be similar; the main difference is that the transitory shock leads to a lower average MPC in steady state.
54
of R, but, for the sake of discussion, suppose the support of this process Z
R
is a subset
of R that is disjoint from {z
H
1
, z
H
2
} allowing us to use a single variable to describe the
state.
30
With probability Π
RH
j
the worker transitions to state ζ
it
= z
H
j
where j {1, 2}. If
ζ
it1
{z
H
1
, z
H
2
}, with probability Π
HH
j
they remain a high-earner with ζ
it
= ζ
it1
and with
the remaining probability they draw from ζ
it
N(0, σ
2
ξ
/(1 ρ
2
ξ
)).
The parameters to estimate are χ, the variance of µ
i
, σ
2
ξ
, the levels of high earnings
z
H
1
and z
H
2
, and the transition probabilities {Π
RH
j
, Π
HH
j
}
2
j=1
. We do this by simulating the
income process as in Guvenen et al. (2022). This procedure takes an observed times series
of average earnings as an input and then seeks to match a variety of moments on earnings
dynamics as measured in the Social Security Administration data. We simulate the process
at a quarterly frequency and time aggregate to annual observations. Many of the moments
we target reflect the shape of the earnings growth distribution for earnings growth at 1-year,
3-year, and 5-year horizons. For the purposes of this paper, the main question is how we
identify χ. Here we proceed as follows. For each business cycle episode between 1979 and
2010, Guvenen et al. (2014) construct a figure analogous to Figure B.1, using average income
over the five years prior to the business cycle episode to rank individuals in the distribution.
For each recession and expansion, we fit a trend line between the 11th and 80th percentiles
of Figure B.1 to obtain a target “slope.” We then seek to replicate these slopes in our
simulated process. To get the model-implied analogue, we feed in average earnings as our
measure of (1 α)y
t
, from there recover household earnings, and then finally construct the
model-implied slope. One component of our objective function is then simply the percentage
deviation between the model and empirical slopes of the incidence of the business cycle.
In addition to the Guvenen et al. (2022) moments, we also solve for the steady state wealth
and income distributions (assuming a constant level of aggregate earnings) and evaluate the
moments in Table 1. The estimation procedure seeks to minimize the deviation of all of
these moments from their target values.
The estimated parameters that enter our model are σ
ξ
= 0.064, χ = 2.0, (z
H
1
, z
H
2
) =
(2.9, 24.8) relative to an average ζ of 1,
RH
1
, Π
RH
2
) = (6.2×10
3
, 4.1×10
4
), and
HH
1
, Π
HH
2
) =
(0.91, 0.95).
Monetary policy shocks for internal calibration. Two parameters of our model
are set by internal calibration, to target estimated impulse response functions to identified
30
In our computations we discretize the process for regular workers allowing us to use a single state for
the idiosyncratic shock.
55
Figure B.1: Incidence of earnings losses during the 1979-1983 recession. The horizontal axis ranks
individuals according to their earnings during 1974-1978. For each percentile of this distribution,
the vertical axis shows the change from 1979 to 1983 in the log of average earnings for that group.
The upward slope implies that lower-income individuals suffered larger earnings losses on average in
the recession. Data are taken from Guvenen et al. (2014). The model results are from a simulation
of the fitted income process, feeding in the time series of average (aggregate) earnings growth.
monetary policy shocks: the relative risk aversion γ and the Phillips curve slope κ. We use
the high-frequency monetary shocks identified by Gertler & Karadi (2015) to estimate the
response of key macro aggregates to identified monetary shocks. We first describe the results
and internal calibration procedure and then give the details of the empirical implementation.
Figure B.2 shows point estimates for four estimated impulse response functions. In the
left panel we show the time path of real interest rates, which we construct as i
t
π
t+1
.
As expected, the expansionary monetary shock leads to a persistent decline in real interest
rates. The center panel shows that consumption increases, eventually reaching a peak of
0.9% above steady state. The right panel reveals that inflation rises quite persistently, with
a peak increase of about 0.3%.
Our internal calibration procedure assumes that monetary policy is set according to a
standard Taylor rule subject to shocks. We find a path of current and anticipated monetary
policy shocks that, when announced at date 0, leads to a change in real rate expectations that
56
Figure B.2: Empirical (orange) and model-implied (blue-dashed) impulse response following an
expansionary monetary policy shock.
perfectly replicates the empirical estimate in Figure B.2.
31
We then choose {γ, κ} to match
the peak responses of output and inflation. Figure B.2 shows the model-implied impulse
response functions. For output and inflation, the model generates the correct peak response,
but unsurprisingly does not generate the hump-shaped pattern of the empirical estimates.
Our empirical analysis uses the following series: the 3-month Treasury rate, the inflation
rate constructed from the log difference in the GDP deflator, and log real GDP per capita.
32
Our monetary shock series is equal to the OLS point estimates of monetary shocks as implied
by the Gertler & Karadi SVAR-IV, estimated only for scheduled FOMC meetings. We use
data from 1980Q3 to 2015Q3, and estimate the impulse response functions using a recursive
VAR with the identified shock ordered first (Plagborg-Møller & Wolf, 2021).
B.2 Initial asset positions and revaluation effects
At date 0, news arrives that changes the returns on capital, long-term bonds, and capital
bonds away from the steady state return. To allocate these revaluation effects across house-
holds we need to know their steady state gross positions in these three asset classes. Our
approach is to impute these as functions of household net worth based on data from the 2019
31
By the results in McKay & Wolf (2023), the output of this procedure is independent of the baseline
policy rule, as long as it induces a determinate, sunspot-free equilibrium. Our choice of baseline rule is thus
a purely computational device.
32
In Figure B.2 we plot W log(P W/P Y ) + log(Y ), where P W is nominal net worth, P Y is nominal
GDP and Y is real GDP.
57
SCF. Auclert & Rognlie (2018) use a related approach.
To give an overview, for each household in the SCF we do a calculation that is similar
to the one reported in Table 2. We take the “real world” asset classes and map them into
model assets. We then create fine bins of households by net worth and compute average
portfolios at different levels of net worth. We then fit piece-wise polynomials through these
functions of net worth and use these to impute the portfolios of the households in the model
according to their net worth.
In our model, all capital holdings are liquid while in reality consumer durables and
real estate is illiquid. We assume that all households with negative net worth simply have
short-term debt and no long-term debt or capital. As these households tend to be liquidity
constrained in our model, this assumption avoids counterfactually strong spending responses
out of changes in asset values for these households. What follows then applies to households
with positive net worth.
Turning to the details, we use the summary extract of the SCF produced by the Federal
Reserve and the variable names here refer to the variable names there. Define equity =
stocks+stmutf + 0.5 comutf + bus, where we assume combination and other mutual funds
are 50-50 stocks and bonds. Note that equity includes both public and private equity. Define
mortgage = nh mort + resdbt + heloc. Define capital = (nfin bus + 1.32 × equity), where
1.32 is the leverage ratio for equity and nfin bus is real estate and consumer durables
(non-financial assets less business holdings). Define short term bond holdings as short =
liq+cdsccbalinstall0.380.32×equity0.5mortgage). This sums cash-like accounts
less credit card and installment loans. We then subtract off the short-term debt associated
with the leveraged position in capital
33
and half of the mortgage balance. Finally, we define
long = bond+tfbmutf +gbmutf +obmutf +0.5comutf 0.620.32equity0.5mortgage.
We rescale all dollar values so that average household income is 1. We then take averages
within fine bins of net worth. We then perform a piecewise regression of the portfolios on a
quadratic function of net worth. We perform this regression separately for households with
net worth above and below one year’s average income.
Like the Distributional Financial Accounts, we use the Financial Accounts for aggregate
quantities of each asset. Therefore, the total stocks of capital, short-term bonds, and long-
term bonds are the totals from Table 2. Our imputation procedure will not exactly match
these totals both because the SCF does not aggregate to the same totals as the Financial Ac-
33
Every dollar of equity is 1.32 dollars of capital and -0.32 dollars of bonds. Those bonds are 62% long-term
and 38% short term.
58
Figure B.3: Initial response of consumption to a $500 fiscal stimulus payment across the income
distribution. The figure includes general equilibrium effects where monetary policy is assumed to
follow a simple interest rate rule i
t
=
3
2
π
t
+
1
8
y
t
.
counts and because our model-generated distribution of net worth does not perfectly match
the SCF distribution of net worth. We therefore rescale the bond positions to match the
total stocks of these assets. We then set capital as net worth less bond holdings. The result-
ing distribution of capital holdings aggregates to the correct total because in the stationary
equilibrium of the economy the total holdings of assets matches the total from Table 2 and if
we have the correct total bond holdings we residually have the correct total capital holdings.
B.3 Additional optimal policy results
We here collect some figures that supplement the quantitative optimal policy analysis pre-
sented in Section 5.4.
Distributional effects of stimulus checks. Stimulus checks and interest rate cuts
have the same effects on aggregate output and inflation (Wolf, 2021), but they differ substan-
tially in their cross-sectional incidence. Figure 2 in the main text shows the model-implied
incidence profile of interest rate cuts; Figure B.3 here does the same for stimulus checks. As
expected, the incidence profile here is strongly downward-sloping: the consumption of poor
households responds the most, reflecting their high MPCs and low overall income.
59
Figure B.4: Simulation of income distribution shock using ad hoc policy rule that stabilizes the
consumption of households with low wealth and low income. The bottom panel shows the con-
sumption of wealthy households increases substantially under this rule leading output and inflation
to exceed their targets.
Stabilizing consumption at the bottom. What happens if monetary policy is used
to stabilize consumption of poor households in the face of our distributional shock m
t
? In
Figure B.4 we plot the implications of following a monetary policy targeting rule that adds
a distributional consideration to the dual mandate targeting rule that seeks to stabilize
the consumption of households with low wealth and low income.
34
As the figure shows,
output and inflation exceed the dual-mandate targets and the consumption of high-wealth
households is substantially above the steady state level.
34
The targeting rule is ˆπ
t
+
1
¯η
(ˆy
t
ˆy
t1
) +
1
2
ˆc
Low
t
= 0, where c
Low
t
is the aggregated consumption of
households with income in the bottom third of the income distribution and wealth below the median.
60
B.4 Comparison with Bhandari et al. (2021)
In this section we explore further the reasons for why our conclusions on optimal Ramsey
monetary policy following a cost-push shock differ from those of Bhandari et al. (2021).
We argue that the differences in policy conclusions are due to differences in the economic
environment. To do so we proceed in two steps. First, we present a model closer to that of
Bhandari et al.. Second, we repeat our optimal policy analysis in that environment.
Environment. We alter our baseline model to arrive at the following alternative environ-
ment, designed to mimic key properties of the model of Bhandari et al.. The changes to the
environment are as follows:
Profits. We now assume that profits are distributed to households as a function of
their histories of idiosyncratic shocks. In particular, a household receives a share of
their profits equal to the steady state wealth share of a household with that history of
shocks. These claims on future profits are not traded.
Other assets. The alternative model omits capital in production and includes only
short-term nominal bonds, so there are no traded long-term assets. As we omit capital,
we reduce the total supply of assets, which we now set equal to the steady-state supply
of bonds in Bhandari et al..
Income process. With fewer assets, our income process—which was calibrated to match
the wealth distribution in our full model—now produces far too many borrowing con-
strained agents. Therefore we use a simpler income process without super-star earners.
Business-cycle incidence. The alternative model omits the unequal incidence of the
aggregate business cycle across levels of labor income, again in keeping with the baseline
model in Bhandari et al..
Results. We begin by showing that, in this alternative model, monetary policy has strong
distributional effects. Mirroring Figure 2 from our main analysis, Figure B.5 displays the ini-
tial consumption response to an expansionary monetary policy shock across levels of wealth.
We see that, in the alternative environment, there is a clear difference in shock incidence
across wealth levels, with low-wealth households increasing consumption strongly and high-
wealth households leaving consumption unchanged. This pattern stands in sharp contrast
to Figure 2, where consumption increases evenly across wealth levels.
61
Figure B.5: Initial response of consumption to an expansionary monetary policy shock across the
distribution of wealth in the alternative model described in Appendix B.4.
Next, Figure B.6 shows the optimal policy response to an inflationary cost-push shock
in the alternative model. Crucially, and mirroring the findings of Bhandari et al., we see
that here short-term interest rates fall on impact following the inflationary shock. Overall,
the policy response is thus notably more accommodative than the dual mandate policy,
with a shallower fall in output and —given our flat Phillips curve—a modest increase in
inflation. In the lower panel we see why the Ramsey policymaker finds such an aggressive
rate cut optimal: the cost push shock has a distinct distributional gradient, with high-wealth
households increasing consumption and low-wealth households decreasing consumption, and
a rate cut can mitigate this distributional gradient relative to the dual-mandate outcome.
62
Figure B.6: Optimal policy response to an inflationary cost-push shock in the alternative model
described in Appendix B.4.
63
C Computational appendix
This appendix provides supplementary information on our computational approach. We
compute sequence-space transition paths using the methodology developed in Auclert et al.
(2021), as discussed further in Appendix C.1. Our computation of the inequality term in
the full Ramsey loss function is described in Appendix C.2.
C.1 General equilibrium transition paths
The baseline constraints (2) in our linear-quadratic policy problem in Section 2 are expressed
in sequence space. To compute the corresponding constraints (21) - (22) for our “HANK”
model, we thus follow the computational techniques of Auclert et al. (2021) to compute the
required sequence-space Jacobian matrices. In particular, our computation of the C
maps
in the augmented HANK “IS curve” (22) leverages the so-called “fake news” algorithm.
For computation of the alternative (but equivalent) constraint formulations (25) - (36)
in impulse response space, we follow McKay & Wolf (2023) and proceed as follows. First,
we close the model with arbitrary policy rules for the two instruments i
i
i and τ
τ
τ
x
, subject
only to the requirement that the two rules induce a unique equilibrium. We then compute
impulse responses of all policy targets to the full menu of contemporaneous and news shocks
to those two policy rules. Now denote the impulse response matrix of some variable x
i
to
shocks to the rule for instrument z
j
by
˜
Θ
x
i
,z
j
, and similarly write
˜
Θ
x
i
q
for responses to
non-policy shocks ε
q
under the (arbitrary) baseline rule. Finally write
˜
Θ
z
j
,z
j
for the impulse
response matrix of the instrument itself. We then define Θ
x
i
,z
j
˜
Θ
x
i
,z
j
˜
Θ
1
z
j
,z
j
and similarly
Θ
x
i
q
˜
Θ
x
i
q
˜
Θ
1
z
j
,z
j
. The results in McKay & Wolf (2023) imply that the resulting impulse
responses are independent of the chosen baseline policy rule.
C.2 Inequality term
The inequality term says that the planner would like to stabilize a very large number of
targets—one for each history ζ. Both for intuition and for computation, it is useful to observe
that these consumption shares will only fluctuate if the inputs to the household’s decision
problem fluctuate. By our discussion of the consumption-savings problem in Section 3.1,
those inputs include total labor income (equal to y
y
y), the return on savings (r
r
r), transfers
(τ
τ
τ
x
and τ
τ
τ
e
), the inequality shock (m
m
m), the initial inflation rate π
0
, and the initial asset
prices q
k
0
and q
b
0
. In this section we leverage this insight to recast the problem of stabilizing
64
consumption shares as one of stabilizing the inputs to the household consumption-savings
decision, thus giving the representation in (37), couched in terms of a small number of
aggregates rather than a distribution of histories.
Reformulating the inequality term. Let x
x
x = (y
y
y, r
r
r,τ
τ
τ
x
,τ
e
τ
e
τ
e
,m
m
m, π
0
, q
k
0
, q
b
0
) be the stacked
sequences of inputs to the household problem. Our goal is to show that there is symmetric
matrix Q such that
X
t=0
β
t
Z
bω
t
(ζ, x
x
x)
2
¯ω(ζ)
dΓ(ζ) =
b
x
x
x
0
Q
b
x
x
x + O(||
b
x
x
x||
3
),
where here we have been explicit that the consumption shares at date t depend on the full
sequences of inputs x
x
x. To arrive at this representation, consider a first-order approximation
to the time-t consumption share of individuals with history ζ:
bω
t
(ζ, x
x
x)
t
(ζ)
b
x
x
x
where the derivative
t
(ζ) will be defined formally below. This yields
bω
t
(ζ
t
,x
x
x)
2
¯ω(ζ
t
)
=
b
x
x
x
0
t
(ζ
t
)
0
t
(ζ
t
)
¯ω(ζ
t
)
| {z }
Q
t
(ζ
t
)
b
x
x
x + O(||
b
x
x
x||
3
).
We then integrate across histories and take the discounted sum across time to arrive at
X
t=0
β
t
Z
bω
t
(ζ
t
,x
x
x)
2
¯ω(ζ
t
)
dΓ(ζ
t
) =
b
x
x
x
0
X
t=0
β
t
Z
Q
t
(ζ
t
)dΓ(ζ
t
)
!
| {z }
Q
b
x
x
x + O(||
b
x
x
x||
3
),
giving the desired representation. We have thus arrived at a representation that almost fits
into our general linear-quadratic set-up of Section 2.1, the sole difference being that the
objective function features a non-diagonal quadratic form. To extend our optimal policy
results to this more general case, we consider the same problem as in Section 2.1, but
replacing the diagonal loss function (1) by the more general (non-diagonal) expression
L
1
2
x
x
x
0
Px
x
x, (C.1)
65
The corresponding necessary and sufficient first-order conditions yield the more general op-
timal policy targeting rule
Θ
0
xz
Px
x
x = 0
and the optimal instrument path
z
z
z
Θ
0
x,z
P Θ
x,z
1
×
Θ
0
x,z
P Θ
x,ε
· ε
ε
ε
Evolution of consumption shares. We now explain the first-order approximation of
consumption shares
bω
t
(ζ
t
)
t
(ζ
t
)x
x
x
that we used above. Let c
t
(ζ
t
,x
x
x) be the consumption in date t after history ζ
t
with the input
sequences given by x
x
x, and similarly let a
t+1
(ζ
t
,x
x
x) be the savings chosen in date t. Also let
ζ
t
be the date-t value of the idiosyncratic state. Using the standard recursive representation
of the household’s problem, we can write these choices in terms policy functions, f and g,
that take as their arguments assets a
t
(ζ
t1
) and the current shock ζ
t
, so we have
c
t
(ζ
t
,x
x
x) = f
t
a
t
(ζ
t1
,x
x
x), ζ
t
,x
x
x
(C.2)
a
t+1
(ζ
t
,x
x
x) = g
t
a
t
(ζ
t1
,x
x
x), ζ
t
,x
x
x
. (C.3)
We now consider a first-order approximation to f and g around x
x
x =
¯
x
x
x:
c
t
(ζ
t
,x
x
x) ¯c(ζ
t
) +
dc
t
(ζ
t
,x
x
x)
dx
x
x
b
x
x
x
a
t+1
(ζ
t
,x
x
x) ¯a(ζ
t
) +
da
t+1
(ζ
t
,
¯
x
x
x)
dx
x
x
b
x
x
x.
The derivatives that appear here are total derivatives with respect to x
x
x, including both the
effect on the policy rule at date t and the effect on assets a
t
(ζ
t1
,x
x
x). The derivatives are
evaluated at the steady-state inputs
¯
x
x
x and the level of assets that an individual with history
ζ
t
would have if the inputs x remained at steady state forever, which we denote by ¯a(ζ
t1
).
To calculate these derivatives, we differentiate (C.2) and (C.3):
dc
t
(ζ
t
,
¯
x
x
x)
dx
x
x
=
f
t
a
t
(ζ
t1
), ζ
t
,
¯
x
x
x)
a
da
t
(ζ
t1
,
¯
x
x
x)
dx
x
x
+
f
t
a
t
(ζ
t1
), ζ
t
,
¯
x
x
x)
∂x
x
x
(C.4)
da
t+1
(ζ
t
,
¯
x
x
x)
dx
x
x
=
g
t
a
t
(ζ
t1
), ζ
t
,
¯
x
x
x)
a
da
t
(ζ
t1
,
¯
x
x
x)
dx
x
x
+
g
t
a
t
(ζ
t1
), ζ
t
,
¯
x
x
x)
∂x
x
x
. (C.5)
66
The partial derivative
f
t(
¯a
t
(ζ
t1
)
t
,
¯
x
x
x
)
a
is the marginal propensity to consume for an individual
with states (¯a
t
(ζ
t1
), ζ
t
) in the stationary equilibrium, and
g
t(
¯a
t
(ζ
t1
)
t
,
¯
x
x
x
)
a
is the marginal
propensity to save. Similarly, the partial derivative
f
t(
¯a
t
(ζ
t1
)
t
,
¯
x
x
x
)
∂x
x
x
is the derivative of the
consumption policy rule with respect to the input sequences for an individual with the same
states, and
g
t(
¯a
t
(ζ
t1
)
t
,
¯
x
x
x
)
∂x
x
x
is the analogous derivative of the savings policy rule.
We discuss below how to compute these derivatives. Given that they have been recovered,
it remains to move from consumption levels to consumption shares:
ω
t
(ζ
t
,x
x
x)
c
t
(ζ
t
,x
x
x)
R
c
t
(ζ
t
,x
x
x)dΓ
t
(ζ)
¯ω
t
(ζ
t
) +
1
¯c
dc
t
(ζ
t
,x
x
x)
dx
x
x
b
x
x
x
Z
¯c(ζ
t
)
¯c
2
dc
t
(ζ
t
,x
x
x)
dx
x
x
dΓ(ζ
t
)
b
x
x
x.
t
(ζ
t
) is then given by
t
(ζ
t
)
1
¯c
dc
t
(ζ
t
,x
x
x)
dx
x
x
Z
¯c(ζ
t
)
¯c
2
dc
t
(ζ
t
,x
x
x)
dx
x
x
dΓ(ζ
t
). (C.6)
Computing Q. To compute
t
(ζ
t
) and so Q, the key challenge is to arrive at the deriva-
tives in (C.4) - (C.5). To do so we begin by simulating a history ζ
t
for t = 0, 1, ..., T in a
stationary equilibrium (i.e. with x
x
x =
¯
x
x
x). At each date along this simulation, we recover
the required partial derivatives as follows. The marginal propensities to consume and save
can be computed by standard methods. For the derivatives of the policy rules, we use the
fact that the derivatives with respect to past prices are zero and the derivatives with respect
to current and future prices only depends on the number of periods until the price change
occurs. This allows us to compute all the derivatives by perturbing prices at a single date
and iterating backwards in time using a single loop from T to 0 (see Auclert et al., 2021).
With the partial derivatives in hand, we then construct
dc
t
(ζ
t
,
¯
x
x
x)
dx
x
x
and
da
t+1
(ζ
t
,
¯
x
x
x)
dx
x
x
by iterating
(C.4)-(C.5) forward starting with
da
0
(ζ
1
,
¯
x
x
x)
dx
x
x
given by the change in the households assets with
the initial asset prices (the revaluation effects). Given those derivatives, we can recover
t
(ζ)
and thus get Q
t
(ζ
t
) as well as Q itself.
Impulse response representation. Using (C.4), (C.5), and (C.6), we can write ω
t
(ζ)
as a linear function of the aggregate variables contained in x
x
x. Since each ω
t
(ζ) is linearly
related to x
x
x, and since we can recover the entries of x
x
x as a function of (y
y
y, π
π
π, i
i
i,τ
τ
τ
x
,m
m
m) (by
Lemma 1), we recover the impulse response representation in (35).
67
D Proofs and auxiliary lemmas
D.1 Proof of Lemma 1
We begin the proof by re-stating and slightly simplifying the definition of an equilibrium in
Definition 1. We first repeat the Phillips curve in stacked form:
Π
π
b
π
π
π = Π
y
b
y
y
y + ψη
η
η, (D.1)
where
Π
π
=
1 β 0 · · ·
0 1 β · · ·
0 0 1 · · ·
.
.
.
.
.
.
.
.
.
.
.
.
, Π
y
= κI.
Turning to the demand side, we re-write (18) and goods market clearing as
b
y
y
y =
¯c
¯y
h
(C
y
+ C
Ψ
H
y
e
0
1
)
b
y
y
y + C
r
b
r
r
r + C
x
b
τ
τ
τ
x
+ C
e
b
τ
τ
τ
e
+ C
m
b
m
m
m + C
Ψ
H
π
e
0
1
b
π
π
π + H
b
e
0
1
b
q
q
q
b
+ H
k
e
0
1
b
q
q
q
k
i
,
(D.2)
where e
1
= (1, 0, 0, . . . )
0
and we have used that by
t
= (¯c/¯y)bc
t
. Next, using (14), (15), and
(16), we write the relationships between asset prices and rates of return as
b
r
r
r = r(
b
i
i
i,
b
π
π
π) (D.3)
b
q
q
q
b
= q
b
(
b
π
π
π,
b
r
r
r) (D.4)
b
q
q
q
k
= q
k
(
b
y
y
y,
b
r
r
r) (D.5)
Finally, we combine the government budget constraint (19) and the law of motion for gov-
ernment debt (20) and solve for τ
e,t
to obtain
b
τ
τ
τ
e
= τ
e
(
b
y
y
y,
b
τ
τ
τ
x
,
b
π
π
π,
b
r
r
r). (D.6)
Our first auxiliary result is that, given the shocks (m
m
m,η
η
η) and policy choices (i
i
i,τ
τ
τ
x
), a list
(y
y
y, r
r
r,τ
τ
τ
e
,π
π
π, q
q
q) is part of an equilibrium if and only if (D.1) - (D.6) hold. To show this we need
to check the conditions of Definition 1: requirement 1 and goods market clearing hold by the
construction of (D.2); requirement 3 is re-stated in (D.1); requirement 6 is imposed by (D.6);
requirement 5 is imposed by (D.3)-(D.5); and asset market clearing holds by Walras’ Law.
68
Finally, to complete the equilibrium let bw and
b
` be determined by (10) and the linearized
version of (13) so requirements 2 and 4 hold by construction.
We now simplify this characterization of equilibria further to arrive at Lemma 1. The
key step in the argument is to solve out the asset-pricing and government financing rules and
plug them into (D.2). First, given
b
π
π
π and
b
i
i
i, we can recover
b
r
r
r from (D.3). We can recover
b
q
q
q
b
from (D.4) and
b
q
q
q
k
from (D.5). Second, given
b
y
y
y,
b
τ
τ
τ
x
,
b
π
π
π and
b
q
q
q
b
, we can recover
b
τ
τ
τ
e
from (D.6).
We can thus write
b
y
y
y =
¯c
¯y
h
(C
y
+ C
Ψ
H
y
e
0
1
)
b
y
y
y + C
r
r(
b
i
i
i,
b
π
π
π) + C
x
b
τ
τ
τ
x
+ C
e
τ
e
(
b
y
y
y,
b
τ
τ
τ
x
,
b
π
π
π,
b
i
i
i)
+C
m
b
m
m
m + C
Ψ
H
π
e
0
1
b
π
π
π + H
b
e
0
1
q
b
(
b
π
π
π,
b
i
i
i) + H
k
e
0
1
q
k
(
b
y
y
y,
b
i
i
i,
b
π
π
π)
i
,
and so
b
y
y
y =
¯c
¯y
C
y
+ C
Ψ
H
y
e
0
1
+ H
k
e
0
1
Q
k
y
+ C
e
T
y
| {z }
˜
C
y
b
y
y
y +
¯c
¯y
C
r
R
π
+ C
e
T
π
+ C
Ψ
H
π
e
0
1
+ H
b
e
0
1
Q
b
π
+ H
k
e
0
1
Q
k
π

| {z }
˜
C
π
b
π
π
π
(D.7)
+
¯c
¯y
C
r
R
i
+ C
e
T
i
+ C
Ψ
H
b
e
0
1
Q
b
i
+ H
k
e
0
1
Q
k
i

| {z }
˜
C
i
b
i
i
i +
¯c
¯y
[C
x
+ C
e
T
x
]
| {z }
˜
C
x
b
τ
τ
τ
x
+
¯c
¯y
C
m
m
m
m
where T
, R
, Q
b
, and Q
k
are derivative matrices for
b
r
r
r(),
b
τ
τ
τ
e
(),
b
q
q
q
b
(), and
b
q
q
q
k
(). (D.7)
embeds (14), (15), (16), and (D.6). We have thus reduced the equilibrium characterization
from statements about (y
y
y, r
r
r,τ
τ
τ
e
,π
π
π, q
q
q
b
,q
q
q
k
) to statements about (y
y
y, π
π
π), establishing the claim.
D.2 Proof of Proposition 1
In light of Lemma 1, we can re-state the optimal policy problem as minimizing (24) subject
to the two constraints (D.1) and (D.7). This problem gives the following necessary and
sufficient first-order conditions:
λ
π
W
b
π
π
π + Π
0
π
Wϕ
ϕ
ϕ
π
˜
C
0
π
Wϕ
ϕ
ϕ
y
= 0
0
0 (D.8)
λ
y
W
b
y
y
y Π
0
y
Wϕ
ϕ
ϕ
π
+ (I
˜
C
0
y
)Wϕ
ϕ
ϕ
y
= 0
0
0 (D.9)
˜
C
0
i
Wϕ
ϕ
ϕ
y
= 0
0
0, (D.10)
69
where ϕ
ϕ
ϕ
π
and ϕ
ϕ
ϕ
y
are sequences of Lagrange multipliers on the two constraints. The proof of
Proposition 1 proceeds by guessing (and then verifying) that ϕ
ϕ
ϕ
y
= 0
0
0. Under this assumption,
we can combine (D.8) - (D.9) to get
λ
π
b
π
π
π + λ
y
W
1
Π
0
π
0
y
)
1
W
b
y
y
y = 0
0
0
It is straightforward to verify that this can be re-written as
λ
π
b
π
π
π +
λ
y
κ
1 0 0 . . .
1 1 0 . . .
0 1 1 . . .
.
.
.
.
.
.
.
.
.
.
.
.
b
y
y
y = 0
0
0 (D.11)
But this is just (28), with the conclusion following for t = 0 since by
1
= 0 (as the economy
starts from steady state). It now remains to verify the guess that ϕ
ϕ
ϕ
y
= 0
0
0. For this, consider
some arbitrary (m
m
m,η
η
η), and let (
b
y
y
y
,
b
π
π
π
) denote the solution of the system (D.1) and (D.11)
given (m
m
m,η
η
η). It remains to show that condition (27) is precisely sufficient to ensure that we
can always find
b
i
i
i such that (D.7) holds.
Plugging into (D.7) and re-arranging:
b
y
y
y
˜
C
y
b
y
y
y
˜
C
π
b
π
π
π
C
m
m
m
m
| {z }
demand target
=
˜
C
i
b
i
i
i. (D.12)
Note that the left-hand side of (D.12) is an excess demand term: supply
b
y
y
y
vs. demand
˜
C
y
b
y
y
y
+
˜
C
π
b
π
π
π
+ C
m
m
m
m. At this point, as long as
˜
C
i
is invertible we could stop and say that
whatever the demand target, there is a
b
i
i
i that satisfies (D.12). Invertibility of
˜
C
i
is too
restrictive, however, as we can say that the demand target will satisfy certain conditions so
˜
C
i
b
i
i
i does not need to be able to generate any possible sequence.
By construction of the
e
C function,
e
C
π
,
e
C
i
, and
e
C
m
map into zero-NPV sequences and
e
C
y
preserves the NPV of
b
y
y
y. So equation (D.12) has a particular structure: both sides are
zero-NPV sequences. Provided that all zero-NPV sequences are in the image of
e
C
i
there
exists a
b
i
i
i that satisfies (D.12).
70
D.3 Proof of Proposition 2
Let U
t
denote the time-t flow utility of the Ramsey planner. To derive the second-order
approximation to the social welfare function, it is convenient to begin by writing U
t
in terms
of log deviations of c
t
and `
t
from steady state:
U
t
=
Z
ϕ(ζ)
¯ce
bc
t
ω
t
(ζ)
1γ
1
1 γ
dΓ(ζ) ν
¯
`e
b
`
t
. (D.13)
Our objective is to construct a second-order approximation of (D.13). Similar to the analysis
in Woodford (2003), our strategy is to consider an efficient steady state, allowing evalua-
tion of Equation (D.13) to second order using only a first-order approximation of aggregate
equilibrium dynamics.
Optimality of the steady state requires that the weighted marginal utility of consumption
is equalized across histories:
ϕ(ζ)¯c
1γ
¯ω(ζ)
γ
= ¯u
c
¯c ζ
for some constant ¯u
c
. Rearranging, we can write this as
ϕ(ζ)
1
= ¯c¯ω(ζ)¯u
1
c
ζ
Furthermore imposing that consumption shares integrate to 1 yields
Z
ϕ(ζ)
1
dΓ(ζ) = ¯c¯u
1
c
. (D.14)
Combining the previous two equations, we can recover consumption shares as a function of
planner weights:
¯ω(ζ) =
ϕ(ζ)
1
R
ϕ(ζ)
1
dΓ(ζ)
ζ
For future reference it will furthermore be useful to define
Ξ
Z
ϕ(ζ)
1
dΓ(ζ)
γ
= ϕ(ζ)¯ω(ζ)
γ
ζ. (D.15)
With these preliminary definitions out of the way, we can begin constructing the second-
71
order approximation of (D.13). Differentiating U
t
with respect to bc
t
, we find that
U
bc
t
=
Z
ϕ(ζ)(¯c¯ω(ζ))
1γ
dΓ(ζ)
= ¯c
1γ
Ξ
where the second line follows from the definition of Ξ and some algebra. Notice that the
definition of Ξ and (D.14) together imply that Ξ = ¯u
c
/¯c
γ
so Ξ is the ratio of the (common)
marginal utility of consumption as evaluated by the planner and the marginal utility of
aggregate consumption used by the labor union to value income gains. Next we have
U
b
`
t
= ν
`
(
¯
`)
¯
`.
The total level of production in steady state is efficient by virtue of the assumed pro-
duction subsidy. To see this, first note that the planner’s problem is to choose c and ` to
maximizing welfare subject to ¯c = Ak
α
¯
`
1α
δk. This leads to the efficiency condition
¯c
γ
Ξ (1 α) Ak
α
¯
`
α
= ν
`
(
¯
`).
For future reference, it is useful to re-express this as
¯c
1γ
Ξ = (1 α)
1
¯c
¯y
ν
`
(
¯
`)
¯
`. (D.16)
Turning to the decentralized equilibrium, it follows from the steady state versions of equations
(13), (10), and (A.3)-(A.5) that Ξ¯c
γ
= ν
`
. That the same efficiency condition holds.
Differentiating U
t
with respect to the consumption shares ω
t
(ζ) we have
U
ω
t
(ζ)
= ϕ(ζ)¯c
1γ
¯ω(ζ)
γ
dΓ(ζ)
= ¯c
1γ
ΞdΓ(ζ)
Turning to second order terms, we begin again with the total level and cross-sectional
split of consumption. We find
2
U
t
bc
2
t
= (1 γ¯c
1γ
72
2
U
t
ω
t
(ζ)
2
= γ¯c
1γ
Ξ
¯ω(ζ)
dΓ(ζ)
2
U
t
bc
t
ω
t
(ζ)
= (1 γ¯c
1γ
dΓ(ζ)
For hours worked we have
2
U
b
`
2
t
= ν
``
(
¯
`)
¯
`
2
ν
`
(
¯
`)
¯
`
We can now put everything together, giving the following second-order approximation of
time-t planner utility (D.13):
U
t
¯
U + ¯c
1γ
Ξbc
t
ν
`
(
¯
`)
¯
`
b
`
t
+
1
2
(1 γ¯c
1γ
bc
2
t
1
2
ν
``
(
¯
`)
¯
`
2
+ ν
`
(
¯
`)
¯
`
b
`
2
t
1
2
γ¯c
1γ
Ξ
Z
bω(ζ)
2
¯ω(ζ)
dΓ(ζ)
c
1γ
Ξ
Z
bω
t
(ζ)dΓ(ζ) + (1 γ)¯c
1γ
Ξbc
t
Z
bω
t
(ζ)dΓ(ζ)
Since consumption shares integrate to 1, it follows that
R
bω
t
(ζ)dΓ(ζ) = 0, and so all terms
in the last row are zero. We now wish to evaluate the remaining terms to second order. To
begin, note that the resource constraint and production function give
bc
t
=
¯y
¯c
by
t
=
¯y
¯c
h
(1 α)
b
`
t
b
d
t
i
(1 α)
1
¯c
¯y
bc
t
+
b
d
t
=
b
`
t
.
Substituting this in for
b
`
t
everywhere we have
U
t
¯
U + ¯c
1γ
Ξbc
t
ν
`
(
¯
`)
¯
`(1 α)
1
¯c
¯y
bc
t
+
b
d
t
+
1
2
(1 γ¯c
1γ
bc
2
t
1
2
(φ + 1) ν
`
(
¯
`)
¯
`(1 α)
2
¯c
¯y
bc
t
+
b
d
t
2
1
2
γ¯c
1γ
Ξ
Z
bω(ζ)
2
¯ω(ζ)
dΓ(ζ)
where we have used the definition of φ. To simplify this expression further, impose the
aggregate resource constraint bc
t
=
¯y
¯c
by
t
, use (D.16), and finally note that all higher-order
73
price dispersion terms can be ignored to second order. We thus get
U
t
¯
U + ν
`
(
¯
`)
¯
`(1 α)
1
b
d
t
+
1
2
(1 α)
1
h
(1 γ)(1 α)
¯y
¯c
(φ + 1)
i
by
2
t
1
2
γ
¯c
¯y
Z
bω(ζ)
2
¯ω(ζ)
dΓ(ζ)
The last step in the derivation is to express
b
d
t
in terms of the history of inflation, closely
following the arguments in Woodford (2003). The dispersion term is defined as
35
d
t
Z
p
jt
p
t
η
t
dj
=
Z
e
bp
jt
p
t
η
t
dj,
where we have defined bp
jt
as the log of p
jt
. Taking a second-order approximation around
bp
jt
= ¯p
t
E
j
[log p
jt
] and η
t
= ¯η yields
b
d
t
Z
¯η(bp
jt
¯p
t
) +
1
2
¯η
2
(bp
jt
¯p
t
)
2
2(bp
jt
¯p
t
)bη
t
dj
=
¯η
2
2
Var
j
[bp
jt
]
where we have simplified using that fact that, at our expansion point, there is no dispersion
in p
jt
, so e
bp
jt
= ¯p
t
j. Next we use the Calvo structure to rewrite the definition of d
t
as
d
t
= θ
Z
p
jt1
p
t1
η
t
dj (1 + π
t
)
η
t
+ (1 θ)
1
1 θ
θ
1 θ
(1 + π
t
)
η
t
1
η
t
/(η
t
1)
A second-order approximation of this expression (around a zero-inflation steady state) yields
b
d
t
θ
¯η
2
2
Var
j
[bp
jt1
] +
θ¯η
2(1 θ)
bπ
2
t
θ
b
d
t1
+
θ¯η
2(1 θ)
bπ
2
t
.
Solving backwards:
b
d
t
θ
t+1
b
d
1
+
θ¯η
2(1 θ)
t
X
s=0
θ
ts
bπ
2
s
(D.17)
35
The aggregate production function is obtained from integrating (A.1) across j and using (13), the fact
that all firms choose the same capital-labor ratio, and labor market clearing.
74
We can now return to the problem of the planner. Using our results so far, we can write
planner preferences as
X
t=0
β
t
U
t
ν
`
(
¯
`)
¯
`
1 α
X
t=0
β
t
b
d
t
+
1
2
(γ 1)
¯y
¯c
+
φ + 1
1 α
by
2
t
+
γ
2
¯c
¯y
Z
bω(ζ
t
)
2
¯ω(ζ
t
)
dΓ(ζ
t
)
Note that bπ
2
t
affects
b
d
t
, β
b
d
t+1
, · · · by
θ¯η
2(1θ)
(1, βθ, · · · ) so the discounted sum is
θ¯η
2(1θ)(1βθ)
.
Using this we have
X
t=0
β
t
U
t
ν
`
(
¯
`)
¯
`
1 α
X
t=0
β
t
θ¯η
2(1 θ)(1 βθ)
bπ
2
t
+
1
2
(γ 1)
¯y
¯c
+
φ + 1
1 α
by
2
t
+
γ
2
¯c
¯y
Z
bω(ζ
t
)
2
¯ω(ζ
t
)
dΓ(ζ
t
)
=
ν
`
(
¯
`)
¯
¯η
2(1 α)(1 θ)(1 βθ)
×
X
t=0
β
t
bπ
2
t
+
˜κ
¯η
by
2
t
+
(1 θ)(1 βθ)γ
¯ηθ
¯c
¯y
Z
bω(ζ
t
)
2
¯ω(ζ
t
)
dΓ(ζ
t
)
(D.18)
where
˜κ
(γ 1)
¯y
¯c
+
φ + 1
1 α
(1 θ)(1 βθ)
θ
.
D.4 Auxiliary lemma for Proposition A.1
We here establish that, under Assumptions A.1 to A.3, changes in nominal interest rates
do not affect the distribution of consumption shares. We proceed in two steps. First, we
show that consumption dispersion is unaffected by changes in real interest rates and output
that satisfy particular conditions. Second, we establish that changes in the monetary policy
stance induce changes in real interest rates and output that satisfy precisely those conditions.
Throughout, our arguments closely follow Werning (2015).
Lemma D.1. Suppose that Assumptions A.1 to A.3 hold, and consider paths (r
r
r,y
y
y, m
m
m,τ
τ
τ
x
)
such that
b
m
m
m = τ
τ
τ
x
= 0
0
0 and, for all t = 0, 1, 2, . . . ,
y
1
t
=
˜
β(1 + r
t+1
)y
1
t+1
, (D.19)
where
˜
β (1 + ¯r)
1
. Then the distribution of consumption shares remains constant at its
75
steady state distribution: ω
t
(ζ) = ¯ω(ζ) for all t and ζ.
Proof. We will rewrite the household budget constraint using several substitutions. First,
there are no transfers and no government debt to service so the government budget implies
τ
y
= 0. Second, we use the acyclical earnings risk from Assumption A.1. Third, as capital
is the only asset traded, the value of assets at the start of period t is a
i,t
=
αy
t
/k + q
k
t
k
i,t
where k
i,t
is the units of capital purchased by the household at date t 1. Similarly, the
value of assets at the end of period t is a
i,t+1
/(1 + r) = q
k
t
k
i,t+1
. Putting the pieces together,
the household budget constraint (8) becomes
q
k
t
k
i,t+1
+ c
it
=
αy
t
/k + q
k
t
k
i,t
+
˜
Φ(ζ
it
, ¯m)(1 α)y
t
. (D.20)
Combining the asset-pricing relationship (16) and the aggregate Euler equation (D.19),
we have
q
k
t
y
t
=
˜
β
α
k
+
q
k
t+1
y
t+1
.
Provided that steady state real rates are positive, we have that
˜
β < 1 and so we can solve
forward to find q
k
t
/y
t
=
˜
βα/
h
k(1
˜
β)
i
. In steady state, the return on capital implies
¯r¯q
k
= α¯y/k so we have (using
˜
β = 1/(1 + ¯r))
q
k
t
=
¯q
k
¯y
y
t
.
The household budget constraint becomes
y
t
¯y
¯q
k
k
i,t+1
+ c
it
= (1 + ¯r) ¯q
k
y
t
¯y
k
i,t
+
˜
Φ(ζ
it
, ¯m)(1 α)¯y
y
t
¯y
. (D.21)
Letting z denote an arbitrary realization of ζ
it
, we will now re-state the household
consumption-savings problem in recursive form. We have
V
t
(k, z) = max
k
0
0
log
h
(1 + ¯r)k¯q
k
+
˜
Φ(z, ¯m)(1 α)¯y ¯q
k
k
0
i
+ log
y
t
¯y
+ βE [V
t+1
(k
0
, z
0
)]
If V
t+1
is of the form V
t+1
(k, z) =
˜
V (k, z) + B
t+1
for some sequence B
t+1
, then the Bellman
equation above can be written as
V
t
(k, z) = max
k
0
0
log
h
(1 + ¯r)k¯q
k
+
˜
Φ(z, ¯m)(1 α)¯y ¯q
k
k
0
i
+ βE
h
˜
V (k
0
, z
0
)
i
+ B
t
, (D.22)
76
where B
t
= log(y
t
/¯y) + βB
t+1
. As there is no time-varying aggregate variable apart from
B
t
, V
t
satisfies the same functional form as V
t+1
. By induction, all previous value functions
satisfy this form. Using the steady-state value function to start the induction (i.e., we start
at the steady-state value function
˜
V (k, z) and B
t
= 0), we can conclude from (D.22) that
the optimal decision rule for k
0
as a function of (k, z) will be constant and equal to the
steady state decision rule. This constant decision rule and a stable process for the evolution
of z
0
implies the distribution of (k
0
, z
0
) is stationary. It follows from (D.21) that the optimal
consumption decision rule will scale with y
t
= c
t
. This scaling implies consumption shares
are constant and equal to their steady state values.
As a final step, it remains to relate this recursive formulation of the household decision
problem to the histories of idiosyncratic events. To this end, note that we can write the
consumption share as a function of the state variables associated with that history:
ω
t
(ζ) =
c(k(ζ), z(ζ))
¯c
, (D.23)
where c(k, z) is the steady state consumption function, k(ζ) is the steady state capital
holdings of a household with history ζ and z(ζ) is the most recent event in the history ζ.
(D.23) holds for any paths (r
r
r,y
y
y, m
m
m,τ
τ
τ
x
) such that (D.19) holds and
b
m
m
m = τ
τ
τ
x
= 0
0
0.
D.5 Proof of Proposition A.1
It remains to show that changes in nominal interest rates induce paths of r
r
r and y
y
y that
satisfy (D.19) (since by linearity we already have
b
m
m
m = τ
τ
τ
x
= 0
0
0). But this follows directly
from Werning (2015), as our model economy with Assumptions A.1 to A.3 satisfies the
conditions of his result (i.e., acyclical risk and acyclical liquidity). We furthermore note that
our special case is isomorphic to the incomplete markets model that appears in Section IIIB
of Farhi & Werning (2019).
36
We refer the reader to Werning (2015) for the formal proof.
36
The model in Farhi & Werning (2019) specifies a particular AR(1) process for idiosyncratic income risk
for the sake of computing numerical solutions. We leave the process more general. The important aspect is
that risk is not affected by monetary policy (see Werning, 2015).
77
D.6 Proof of Corollary A.1
By (4) we can write the optimal monetary policy rule as
Θ
0
π,i
W
b
π
π
π +
κ
¯η
Θ
0
y,i
W
b
y
y
y = 0
0
0 (D.24)
where we have used the fact that with δ = 0 we have ˜κ = κ. It follows from (D.1) that
Π
π
Θ
π,i
= Π
y
Θ
y,i
(D.25)
and so we can re-write (D.24) as
b
π
π
π +
κ
¯η
1
κ
1 0 0 . . .
1 1 0 . . .
0 1 1 . . .
.
.
.
.
.
.
.
.
.
.
.
.
b
y
y
y = 0
0
0 (D.26)
giving (A.9).
D.7 Proof of Corollary A.2
We have already shown that the optimal monetary rule is given as (A.9). Since (D.1) also
implies that
Π
π
Θ
π
x
= Π
y
Θ
y
x
(D.27)
we can use the same steps as in the proof of Corollary A.1 to rewrite the first two terms in
(38) as
Θ
0
π
x
W
b
π
π
π +
κ
¯η
Θ
0
y
x
W
b
y
y
y =
b
π
π
π +
1
¯η
1 0 0 . . .
1 1 0 . . .
0 1 1 . . .
.
.
.
.
.
.
.
.
.
.
.
.
b
y
y
y.
Imposing (A.9) sets these terms to zero, so Corollary A.2 follows.
78