1. Introduction
We consider a stochastic control problem in which the Hamilton-Jacobi-Bellman (HJB) admits a two-threshold solution. As a specific setting, we refer to the subscription economy. Households (consumers) make decisions to subscribe to (enroll), or unsubscribe from (cancel), a given subscription service as the flow benefit from the service fluctuates over time. There are costs from transitioning from state “subscribed" to state “unsubscribed" (cancellation cost) and vice versa (enrollment cost), resulting in a two-threshold optimal strategy, i.e. a household subscribes to the service when the flow benefit reaches a threshold value , and unsubscribes when the flow benefit value drops below a threshold value with .
This paper emphasizes the heterogeneity of the market assuming a distribution of thresholds
. We show that the heterogeneous setting leads to an operator, which is well-known in the engineering literature as the Preisach operator, a fundamental model of hysteresis effects, see [
7] and
Section 2.5 for references and a brief review of applications.
The Preisach operator was previously used as a phenomenological model of hysteresis in unemployment [
8] and macroeconomic flows [
9]. In the phenomenological approach, the Preisach operator is postulated as a model of the relationship between economic variables. In particular, the model parameters are not linked to parameters of the market but rather are determined from a black box identification procedure. On the other hand, in this paper, the parameters of the Preisach operator derive from the solution of the optimization problem and, as such, are functions (either explicit or amenable to approximation) of the subscription cost, enrollment/cancellation costs, discount rates and parameters of dynamics of the utility flow.
The subscription economy has grown substantially over the past years. As of 2023, U.S. households spent an average of
$230 per month on subscription services [
25]. Such costs span mobile phone, internet, TV streaming, Amazon Prime, and music streaming services, among others. Importantly, many of these services are easy to subscribe to but are rather cumbersome to unsubscribe from. In fact, in October 2024, the Federal Trade Commission (FTC) announced a final “click-to-cancel” rule that will require sellers to make it as easy for consumers to cancel their enrollment as it was to sign up [
26].
Motivated by the growth of this market and its unique regulatory features, this paper develops a mathematical framework to help us understand the main economic forces driving the subscription services market. We model this market as a Stackelberg game. On the demand side, we characterize the optimal enrollment and cancellation decisions of households (i.e., the followers) as a sequence of compound real options (i.e., the decision to unsubscribe when currently subscribed, and vice versa). We assume that the household takes as given the pricing structure of the service and model the utility flow derived by the household as a diffusion process. Next, we study the firm’s (i.e., the leader’s) problem, in which the firm, anticipating the household’s optimal response, chooses the optimal cost structure for its services. Finally, relying on singular perturbation examples, we provide a detailed characterization of the game’s equilibrium strategies when enrollment/cancellation costs are small relative to the total subscription costs. The justification for small optimal enrollment/cancellation costs is application dependent. However, consider for instance, the case of Amazon Prime. As of 2025, the monthly cost of an Amazon Prime subscription stands at $14.99 a month. Assuming a real discount rate of 3% per annum and a 5-year subscription period, yields a net present cost of $846. By contrast, plausible enrollment/cancellation costs are likely to be one to two orders of magnitude smaller. That is, of the order of $10, and therefore “small" relative to the total subscription costs.
The paper is organized as follows. In
Section 2, we discuss the model, the Bellman equation (in the form of a variational inequality) for the household’s value function, the two-threshold optimal solution, the firm’s problem in a heterogeneous market and its relationship to the Preisach operator.
Section 3 presents main result. First, the household’s problem is solved in the case of cost-free enrollment/cancellation. The solution provides the benchmark for further results. Next, the existence of a two-threshold solution is shown in the case of non-zero enrollment/cancellation costs. This solution is not explicit. The singular perturbation method is used to approximate the thresholds of the two-threshold solution assuming that the enrollment/cancellation costs are small compared to the total subscription cost. The perturbation expansions for the thresholds are then used to approximate the solution of the firm’s problem. We present a case study of the market with one, two and
N customers.
Section 4,
Section 5 summerize a few implications of the results in the context of subscription economy.
2. Model
2.1. Dynamics
Let
denote a complete probability space that supports a standard Brownian motion
with its natural filtration
. Denote the state variable by
X and suppose that its dynamics are given by
We interpret
as the flow certainty equivalent utility obtained by the consumer if currently subscribed to the service. The utility from the service can vary over time, depending on the effort put by the firm to provide a higher quality product, but it is also subject to taste shocks faced by the consumer and by the quality of other products.
The second state variable is , where 1 means the consumer is currently subscribed to the service, and 0 means the consumer is currently not subscribed to (unsubscribed from) the service. There are costs from transitioning across these states. Denote by the enrollment cost (i.e., of transitioning from 0 to 1). Similarly, denote by the cancellation cost (i.e., of transitioning from 1 to 0). These costs can be the monetary cost of effort from taking each action, but also some monetary transfer offered by the firm to the consumer. We abstract away from this distinction for now.
Using an increasing sequence of times
denote by
the corresponding enrollment/cancellation policy
where
is the initial state, i.e.
is the sequence of transitioning times for
. We denote by
the set of times at which the consumer transitions from 0 to 1, respectively
are the times when the consumer transitions from 1 to 0. Hence,
is the the odd-indexed subsequence and
is the even-indexed subsequence of
if
; conversely,
is the the even-indexed subsequence and
is the odd-indexed subsequence of
if
.
Denote by
the flow cost of the subscription. The expected pay-off for an exponential discounter with discount rate
r from an enrollment/cancellation policy
is given by
with
. The consumer’s objective is to maximize the expected pay-off using an admissible enrollment/cancellation policy as a control. Hence, the value function of this optimization problem is
where
is the set of enrollment/cancellation policies
induced by transitioning time sequences
.
Since
is a stationary Markov process, one expects optimal transitioning times to have the form of recursively as the stopping times
where the closed set
and its complement
are the so-called stopping and continuation regions, respectively, associated with the state
[
6]. In other words, the consumer transitions from state
z to state
at the nearest moment when the process
reaches the stopping region
. In the simplest case,
is an interval with its end point(s) serving as threshold(s), i.e. a transition occurs when
attains a threshold.
In what follows, the enrollment/cancellation costs are assumed to satisfy
Hence, there is a fine for cancellation and an incentive to enroll. The fine exceeds the incentive eliminaing the arbitrage opportunity when the flow benefit
oscillates around the zero value, as well as eliminating excessive transitioning by spreading the thresholds.
2.2. Variational Inequality
In order to solve for the optimal strategy and pay-off, we represent the value function
of a consumer as a pair of value functions
corresponding to the states 0 and 1, respectively. Defining the differential operator
associated in the usual way with the process (
1) and the exponential discount rate
r, and following [
6], the value functions
satisfy the following coupled variational ineqaulities (which are a form of the Bellman equation):
and
2.3. Two-Threshold Solution
It is natural to expect [
1,
5] that the solution of the optimal control problem involves two thresholds
at which the consumer unsubcribes if her flow benefit falls below
(when currently subscribed) and subscribes if her flow benefit reaches
(when currently unsubscribed). In other words, the stopping and continuation regions for the state
are
and for the state
they are
For this solution in the continuation region, the HJB variational inequality (
8)–(13) leads to the equations
subject to the set of boundary conditions, which include the value matching conditions
and smooth pasting conditions
at the thresholds
,
; and, the conditions
at zero and infinity. On the other hand, in the stopping region the HJB variational inequality yields
Moreover, the variational inequality requires that
The corresponding transitioning policy is the sequence of stopping times (
5) defined by the simple threshold-based rule
i.e. a transition across the states occurs when
hits either the threshold
or
(depending on the current state) as well as at the initial moment if the initial conditions satisfy either
and
or
and
.
Proposition 1.
Let and let , be a solution of problem (16)–(20) extended to by equations (21)–(22). Then relations (23)–(26) hold iff
Proof. The proof included here for completeness follows in a standard way from the Maximum Principle.
For
, equation (
21) combined with (17) implies
hence (
23) is equivalent to (
28). Similarly, for
, from (22) and (
16) it follows that
therefore (25) is equivalent to (29).
Next, define
Then, (24) is equivalent to
From equations (
16), (17) and the boundary conditions (
18), (
19) at
, it follows that
Therefore, (24) implies (
28). Similarly, using the function
relation (26) is equivalent to
for
, while equations (
16), (17) and the boundary conditions (
18), (
19) at
imply
Hence, (26) implies (29).
Conversely, to show that (
28) implies (
30), which is equivalent to (24), assume that (
28) holds with the strict inequality. Then, (
31) implies the existence of a sufficiently small interval
where
From (22) and (
6) it follows that
Assume for contradiction with (
30) that
Then, by (
33), (
34),
u has a local maximum point
and a local minimum point
such that
As such,
But due to (
16), (17),
hence (
35) contradicts the monotonicity of
. The contradiction proves (24) when the inequality in (
28) is strict and implies (24) in the case of the equality in (
28) by the continuity argument.
Similarly, relations (
21) and (
6) imply
If (29) holds with the strict inequality, then due to (
32),
on a sufficiently small interval
. If
then relations (
36), (
37) imply that
w has has a local maximum point
and a local minimum point
such that
which leads to the contradiction between the relations
and the monotonicity of the function
on the interval
. Hence,
for all
proving (26). The continuity argument completes the proof of the implication (29) ⇒ (26) in the case of the equality in (29). □
The solution to problem (
16)–(
20) will be addressed in
Section 3 under the assumption that the characteristic polynomial
associated with the differential operator
satisfies
Denoting the roots of
L by
with
, this assumption is equivalent to
As a case in point, if
, i.e. the per-period pay-off is linear, and
, then
In the next two subsections, we consider an extension of problem (
16)–(
20).
2.4. Heterogeneous Market and Firm Problem
We consider a heterogeneous market where there are
N consumers with different discount rates
,
, each maximizing her expected pay-off (
4). With each consumer we associate a weight
, where
and denote by
the subscription policy of the
i-th consumer. Next, we extend the model by introducing the service providing firm, thus closing the feedback loop. We assume that the flow benefit of the firm from each subscription is
. Hence, the aggregated flow benefit from all the subscriptions at time
t equals
Moreover, there are costs associated with enrollment/cancellation of consumers, which are assumed to be proportional to
and
. As such, the expected pay-off of the firm is
with
, where
is the firm’s discount rate. The firm sets the flow rate
p and the enrollment/cancellation costs
seeking to maximize its expected pay-off (
43), while each consumer sets her enrollment/cancellation thresholds
,
based on the values of
,
seeking to maximize her pay-off (cf. (
4)). The thresholds
,
determine the enrollment/cancellation policy
, which feeds back to (
43).
The counterparts of relations (
40) are assumed to hold for the roots
of each polynomial (
38) each
,
.
2.5. Relation to Preisach Model
Relations (
2), (
3), (
5) define a map from any space of continuous inputs
(not necessarily realizations of the Geometric Brownian Motion) to the space of binary functions
. In the case when the stopping and continuation regions have the form (
14), (
15) with
, i.e. (
5) has the form (
27), this map is well-known in engineering applications as the two-threshold two-state
non-ideal relay (also known as
bi-stable switch,
‘lazy’ switch,
elementary rectangular hysteresis loop, or
Schmitt trigger depending on a particular application).
Stacking
N non-ideal relays, which all have the same input
, and defining the output
of the system as the aggregated quantity (
42), has a long history in physics and engineering, starting from the fundamental phenomenological model of magnetic hysteresis proposed by F. Preisach [
21]. The states of the system are naturally
N-tuples
, hence the total number of plausible states is at most
(but can be smaller depending on the thresholds, see
Section 3.5). The Preisach operator, which extends this hysteresis model to the continuous setting [
16,
19,
23], has counterparts in many disciplines, most notably the Prandtl-Ishlinskii model in plasticity [
17,
18], Maxwell-slip friction model in tribology [
4] and Parlange model in hydrology [
11]. Further applications of the Preisach model include control based on smart materials [
3,
12,
13,
14,
22], economics [
2,
8,
24], neuroscience [
20] and epidemiology [
10,
15]. The above discussion shows how the same model arises in the setting of real options and pricing.
Remark 1. A more general market model can include a population of customers with different , in their corresponding pay-offs (4). These variations still lead to the Preisach operator in the firm’s pay-off. However, solutions presented below are specific to the particular case where the consumers differ by the discount rate only, see for example Remark 2. The more general case will be considered elsewhere.
3. Results
3.1. Cost-Free Enrollment/Cancellation Benchmark
In what follows, the polynomial
L of each consumer and of the firm are all assumed to satisfy condition (
39), which is equivalent to (
40). These polynomials can have different
r and hence different roots
.
As a benchmark case, in this section we consider zero enrollment/cancellation costs, i.e.
. In this case, relations (
21), (22), (24), (26) imply
on the whole
, hence we simply write
. Hence, from (
16), (17) it follows that
and
where
. Combining these equations with (
23), (25) results in
therefore
Using solutions (
60) of equations (
44), (45) and applying the value matching and smooth pasting conditions at the threshold (
46), one obtains the
smooth value function
with
The corresponding transitioning policy is the simple one-threshold rule
where
is the Heaviside step function
Accordingly, the consumer is subscribed whenever
, unsubscribed whenever
, and transitions from state 0 to 1 and vice versa at the same threshold
specified by (
46). We note that the threshold
doesn’t depend on
r.
Next, we consider the firm problem stated in the previous section. When all the consumers implement the same transitioning policy (
47), the normalization condition (
41) implies that the expected firm’s pay-off (
43) equals
Effectively, the population of consumers acts as one consumer. Therefore, the function
can be obtained as a solution of the differential equation
(cf. (
7)) subject to the boundary conditions
Using the value matching and smooth pasting conditions at the discontinuity point (
46) of the step function gives
where
are the roots of the chracteristic polynomial
satisfying
(cf. (
38), (
40)), and we use the variable
Proposition 2.
The pay-off (51) of the firm achieves its maximum with respect to p at the point
where
and is a unique positive root p of the equation
The function (52) satisfies
Proof. By inspection, for a fixed
,
and each of the functions
,
(as functions of
) has a unique maximum. Namely,
doesn’t depend on
; on the other hand,
is a unique positive root
p of the equation (
53). Therefore, the pay-off achieves its maximum with respect to
p at the point (
52). By inspection,
is equivalent to
, hence
iff
. Also,
, hence if
, then
and consequently
. Therefore, (
54) holds and
which combined with
for
implies (
55). □
The firm maximizes its pay-off by setting
depending on the initial value of the state variable,
, and the corresponding
In particular, if the initial value satisfies
then the firm sets
, i.e. the firm’s expected pay-off is (using the variable
)
Since
, all the consumers with
unsubscribe at the initial moment in accordance with transitioning policy (
47). On the other hand, if
then the firm sets
, and the expected pay-off is
Since due to (
55) from
it follows that
, all the consumers with
subscribe at the initial moment.
Remark 2. If consumers have different flow benefits (i.e. different or ), then they have different thresholds of their transitioning policies according to (47). Hence, the population is no longer represented by one consumer as above, and the solution is more complicated. This more general model will be considered elsewhere.
3.2. Existence of a Two-Threshold Solution
In this section, we prove the existence of a two-threshold solution to the HJB variational inequality (
8)–(13) assuming non-zero enrollment/cancellation rates. Recall that this solution is represented by a solution of system (
16)–(26).
Proposition 3. Let relations (6) and (39) hold. Then, system (16)–(26) has a solution.
Proof. Applying boundary conditions (
18)–(
20) to the general solution of Euler’s equations (
16), (17) and using (
40) results in
with
where the thresholds satisfy the system
According to Proposition 1, it suffices to prove that this system has a solution
satisfying (
28), (29). To show this, we define the functions
with
, and rewrite system (
62), (
63) equivalently as
By inspection, the functions
,
satisfy
Moreover,
has a unique local minimum on the positive semiaxis at the point
and,
has a unique local maximum on the positive semiaxis at the same point
. Therefore, the equation
defines a unique implicit function
Substituting this function into the equation
, one obtains the scalar equation
From
it follows that
On the other hand, since
is the maximum point of the function
and
decreases in the second variable,
hence
and by the intermediate value theorem
has a zero
, which proves the existence of the thresholds. □
3.3. Asymptotic Approximation of the Two-Threshold Solution
We now develop an asymptotic expansion of the two-threshold solution assuming that the enrollment/cancellation costs are small compared to the cumulative subscription cost. More precisely, let us assume that
hence
This dimensionless quantity is used as the small parameter of the asymptotic expansion. The one-threshold solution
of the cost-free enrollment/cancellation case presented in
Section 3.1 serves as a reference corresponding to
. The following statement uses the dimensionless parameters
and
Proposition 4.
The singular asymptotic expansion of the solution to system (62)–(63) satisfying (28)–(29) has the form
where
Proof. The above asymptotic expansions of
are obtained in a standard way by substituting equations (
66) into system (
62)–(
63), expanding the resulting equations with respect to
as
then equating each coefficient
of these expansion to zero and solving the resulting linear system for
for
with
. □
Remark 3. According to Proposition 4, the second order approximations of doesn’t depend on the discount rate r, the third and fourth order approximations are linear in r, and the higher order approximations starting with order five are non-linear with respect to r. The role of this nonlinearity is discussed in Section 4.
3.4. Firm Problem: One Consumer
Next, we revisit the firm problem posed in
Section 2.4 assuming large coefficients
in the firm pay-off (
43). The assumption
warrants
, i.e. small optimal enrollment/cancellation costs, letting us use the asymptotic expansions developed in Proposition 4 with the cost-free benchmark (see
Section 3.1).
We begin with the firm problem with one consumer who implements the two-threshold strategy. The expected pay-off of the firm satisfies the equations
where the subscripts
correspond to the state of the consumer. These equations are combined with the boundary conditions
at zero and infinity, resulting in the relations
(cf. (
49)–(
50)), where the coefficients are determined by the value matching relations at the thresholds:
Recall that the consumer sets the thresholds
to maximize her pay-off depending on flow rate
p and enrollment/cancellation costs
, hence the coefficients
and the firm’s pay-off
are functions of
. As such, the firm’s value function
is defined by
Using the expansions for the thresholds
developed in Proposition 4, one obtains
with
where parameters (
65) and the variable
are used. This expansions lead to the following asymptotics of the firm’s value function and parameters of the optimal strategy.
Proposition 5.
-
(i)
If , then for (cf. (56)–(58)). The firm maximizes its expected pay-off by setting and , i.e. zero enrollment/cancellation costs, which leads to the single-threshold behavior of the consumer with the threshold ; the initial state of the consumer is “unsubscribed".
-
(ii)
If , then the firm’s value function satisfies
for (cf. (59)), where
The firm maximizes its expected pay-off by setting
i.e. zero enrollment cost, while the optimal flow rate and cancellation cost satisfy
leading to the two-threshold behavior of the consumer with the initial state “subscribed". Moreover, setting
results in the asymptotically the same firm’s pay-off as in (83), i.e.
Proof. Since the coefficients of expansion (82) satisfy , , the firm’s pay-off decreases in near in the case . Hence, the firm maximizes its pay-off by setting , which corresponds to , i.e. forgoing both enrollment and cancellation costs as in Proposition 2. This proves (i).
On the other hand,
,
according to (
81), hence the pay-off
increases in
near
. Therefore, in the case
(i.e. under the conditions of part
(ii)),
achieves its maximum at a point where
Since
, this implies
By inspection, for
this expansion combined with
implies
with
defined by (
84). Moreover, differentiating
with respect to
and taking into account that
,
, one obtains
where
Therefore, for sufficiently large
, the firm’s pay-off increases with
, hence
implies that the pay-off is maximized by setting
, which corresponds to
where due to (
90) and
,
Now, we check
a posteriori that the accuracy used for the approximation of
p is sufficient to obtain the expansion (
83) of the value function. Indeed, by inspection, differentiating (
81) with respect to
p, using (
53) and setting
gives
Hence,
implies that the root of the equation
satisfies
. Moreover, the small correction
doesn’t affect the asymptotic expansion (
91) for
obtained above from
using
; neither it affects the estimate
that was established for small
. Therefore, the optimal parameters satisfy (
85)–(
86), and the corresponding value of the pay-off can be obtained with the accuracy
by substituting
in (
81), which (using
) results in (
83) and (
88). Finally, parameters (
92) correspond to (
87). □
Remark 4.
Formulas (89) and (91) combined lead to the next order approximation for ε, namely
which can be used to obtain the next order terms in expansions (83) and (86) of the value function and optimal parameters, respectively.
Remark 5. Proposition 5 doesn’t address initial values from a small interval of length centered at . An optimal strategy aiming at an increase of the pay-off of the order over the cost-free enrollment/cancellation benchmark (cf. (58), (59) and (83)) in this more subtle case will be considered elsewhere.
3.5. Firm Problem: Two Consumers
Continuing the case study, let us consider the firm’s problem with two consumers, each optimizing the enrollment/cancellation thresholds
,
, of her two-threshold policy in response to the flow rate
p and the enrollment/cancellation costs
set by the firm. Consumers’ discount rates satisfy
Using the notation of
Section 2.4, the firm’s flow benefit from the
i-th subscription is
(i.e., when the
i-th consumer’s state is “subscribed"), where
due to (
41). The system has at most 4 states, and the firm’s expected pay-off has the form
where the subscript
means that the first consumer is in state “subscribed", while the second consumer is in state “unsubscribed"; the other subscripts have simila meaning; and, the coefficients are determined by the value matching relations at the thresholds (cf. (
78), (79)). These value matching conditions depend on the ordering of the thresholds. In particular, if either
or
, then there are 3 states,
,
,
, and the value matching conditions have the form
If
, then there are 4 states,
,
,
,
, and the value matching conditions are
Due to
, it follows from equations (
64)–(70) that for any sufficiently small
,
hence there are 3 states,
,
and
.
Let us introduce a consumer with the averaged discount rate
(i.e.
corresponds to
and
corresponds to
), and consider the firm problem with this one consumer (called the averaged consumer), i.e. the problem discussed in
Section 3.4 with
. Denote by
the firm’s pay-off for this problem (cf. (
78)–(
80)). We now compare
with the firm’s pay-off in the market with two consumers.
Proposition 6.
The firm’s pay-off in the market with two consumers, which is defined by (93)–(97), satisfies
where
Proof. Expansions (
99), (100) are obtained by substituting the expansions for the thresholds
developed in Proposition 4 into relations (
78)–(
80) and (
93)–(
97). The coefficients
,
of expansions (
66) don’t depend on consumer’s discount rate
r, the coefficients
,
depend on
r linearly (the parameter
is proportional to
r), and the coefficients
are quadratic in
r (cf. (
67)–(74)). The latter nonlinearity induces the discrepancy between
and
of order
(and between
and
, respectively). We omit further details of this direct computation. □
Remark 6.
Comparing the expressions for the firm’s pay-off, (78)–(80) and (93)–(97), one obtains
An immediate corollary of Propositions 5 and 6 is the following statement.
Proposition 7.
-
(i)
If , then the value function of the firm’s problem with two consumers equals for (cf. (56)–(58)). The firm maximizes its expected pay-off by setting and , while both consumers adopt the single-threshold behavior with the threshold ; the initial state of the consumers is “unsubscribed".
-
(ii)
If , then value function of the firm problem with two consumers satisfies equation (83) with for . The firm maximizes its expected pay-off by setting , while the optimal flow rate and cancellation cost satisfy (86), leading to the two-threshold behavior of both consumers, each adopting the initial state “subscribed". Moreover, parameters (87) ensure that the firm’s pay-off satisfies (88), i.e. it is asymptotically close to the value function.
3.6. Firm Problem with N Consumers
Adopting the notation of
Section 2.1 in the setting of
Section 2.5, the states of the market with
N consumers are
N-tuples
. Let
denote the firm’s expected pay-off corresponding to the state
Z. As such,
,
using the notation of
Section 3.4, and
,
,
using the notation of
Section 3.5. Since the firm’s flow benefit from all the subscriptions is given by (
42), the firm’s expected pay-off is the weighted sum of the expected pay-offs from individual consumers (cf. Remark 6 for the case of two consumers). More precisely, the following statement holds.
Proposition 8.
The firm’s expected pay-off in the market with N consumers equals
where is the firm’s pay-off in the market with one i-th consumer, i.e.
(cf. (78)–(80)) with
where are the consumer’s thresholds.
Proof. By inspection, functions (
102), (103) solve equations (
75), (76), respectively, subject to conditions (
77) at zero and at infinity. Also, relations (
104) warrant that these functions satisfy the value matching conditions at the thresholds
,
of the
i-the consumer (cf. (
80)). Therefore, for every state
Z, function (
101) is the solution of the equation
satisfying the same conditions at zero and infinity. Moreover, the value matching conditions
,
are satisfied at each pair of states
,
such that
and
for
. □
Recall that the Preisach operator maps continuous inputs
to states
(see
Section 2.5). According to Proposition 8, the firm’s expected pay-off is the function of the state defined by equations (
101).
4. Discussion
To summarize a few results obtained above, according to the proposed model, the service provider (firm) increases its expected pay-off by relinquishing any enrollment stimulus for consumers. On the other hand, if the initial consumer’s utility from the service is sufficiently high, then the firm’s optimal strategy involves setting a non-zero cancellation penalty (see Proposition 5).
The value function of the firm operating in a heterogeneous market where consumers have two distinct discount rates,
, can be approximated by the value function of the firm operating in a homogeneous market where all the consumers have the averaged discount rate
defined by (
98) (see Proposition 7). Moreover, Proposition 6 implies that this approximation underestimates the firm’s value function if
and overestimates it if the opposite inequality holds (cf. equation (
99)).
Equation (
81) implies that
Larger discount rate
r is interpreted as higher consumer’s impatience. Hence the firm’s value function is higher in a market of more impatient consumers or a heterogeneous market with a larger proportion of such consumers. Moreover,
therefore the sensitivity of the value function to the controls
p and
is also higher in a market with more impatient consumers.
As an example, for the set of parameters , , , , , , , , the thresholds of the two consumers are , , , , and the relative error of the approximation of order provided by Proposition 4 is . Assuming and , the firm’s expected pay-off increases from to as varies from 1 to 0. The relative error of the approximation provided by Proposition 6 is also .
The main premise for the above conclusions is the assumption that the cancellation cost, , is sufficiently small compared to the expected total subscription cost, .
5. Conclusions
We showed that the solution to a model of a subscription economy with heterogeneous households naturally contains an equivalent of the Preisach operator. In particular, in the case of non-zero enrollment/cancellation costs, the household’s optimal strategy is a two-threshold strategy, and the corresponding firm’s expected pay-off is a weighted sum of the expected pay-offs from individual households. Using the ratio of the enrollment/cancellation cost and the expected total subscription cost as a small parameter, we obtained a perturbation expansion of the optimal solution, which relates the parameters of the Preisach operator (i.e., the set of thresholds) with parameters of the households’ utility flow (such as the distribution of discount rates) and controls (i.e., enrollment/cancellation and flow costs of subscription imposed by the firm). Within the model constraints, we showed that the firm increases its expected pay-off by relinquishing any enrollment stimulus. On the other hand, if the initial households’ flow benefits are sufficiently high, then the firm’s optimal strategy involves a (small) non-zero cancellation penalty. The perturbation expansions show that a heterogeneous market can be approximated, within an accuracy range, by an averaged representative household, i.e. a homogeneous market. However, the Preisach operator provides a more accurate solution. The firm’s pay-off increases with the increasing proportion of impatient households (i.e., having higher discount rate). Moreover, the sensitivity of the value function to the controls is higher in a more impatient market. Our analysis is limited to the heterogeneity due to a distribution of households’ discount rates. Similar analysis can be applied to other types of heterogeneous utility that households derive from a subscription (e.g.,
,
in (
4)).
References
- S. Acharya, A. Bensoussan, D. Rachinskii, A. Rivera, 2021, Real options problem with nonsmooth obstacle, SIAM Journal on Financial Mathematics 12 (4), 1508-1552. [CrossRef]
- J. Adamonis, M. Göcke, 2019, Modeling economic hysteresis losses caused by sunk adjustment costs, J. Post Keynesian Economics, 42, 299-318. [CrossRef]
- V. Apicella, C. S. Clemente, D. Davino, D. Leone, C. Visone, 2019, Review of modeling and control of magnetostrictive actuators Actuators 8 (2), 45.
- F. Al-Bender, V. Lampaert, J. Swevers, 2005, The generalized Maxwell-slip model: A novel model for friction simulation and compensation, IEEE Transactions on Automatic Control 50, 1883-1887. [CrossRef]
- A. Bensoussan, B. Chevalier-Roignant, A. Rivera, 2021, Does performance-sensitive debt mitigate debt overhang?, J. Econom. Dynam. Control 131, 104203. [CrossRef]
- A. Bensoussan, J.-L. Lions, 1982, Applications of Variational Inequalities in Stochastic Control, North-Holland, Amsterdam.
-
The Science of Hysteresis, 2006, G. Bertotti, I. Mayergoyz (Eds.), vol. 1–3, Elsevier.
- R. Cross, J. Darby, J. Ireland, L. Piscitelli, 2006, Hysteresis and unemployment: a preliminary investigation, in G. Bertotti, I. Mayergoyz (Eds.), The Science of Hysteresis, vol. 1, Elsevier, 667-699.
- R. Cross, H. McNamara, L. Kalachev, A. Pokrovskii, 2013, Hysteresis and post Walrasian economics, Discrete and Continuous Dynamical Systems B 18 (2), 377-401. [CrossRef]
- R. Guan, J. Kopfová, D. Rachinskii, 2024, Global stability of SIR model with heterogeneous transmission rate modeled by the Preisach operator, SIAM J. Applied Dynamical Systems 23 (2), 1199-1241. [CrossRef]
- R. Haverkamp, P. Reggiani, P. J. Rossand, J.-Y. Parlange, 2002, Soil water hysteresis prediction model based on theory and geometric scaling, in P. A. C. Raats, D. Smiles, A. W. Warrick (Eds.), Environmental Mechanics: Water Mass and Energy Transfer in the Biosphere, 213-246.
- R. V. Iyer, X. Tan, P. S. Krishnaprasad, 2005, Approximate inversion of the Preisach hysteresis operator with application to control of smart actuators, IEEE Transactions on automatic control 50 (6), 798-810. [CrossRef]
- M. Al Janaideh, S. Rakheja, C. Y. Su, 2010, IEEE/ASME Transactions on Mechatronics 16 (4), 734-744.
- B. Kaltenbacher, P. Krejčí, 2016, A thermodynamically consistent phenomenological model for ferroelectric and ferroelastic hysteresis, ZAMM - Z. Angew. Math. Mech. 96, 874-891. [CrossRef]
- J. Kopfová, P. Nábělková, D. Rachinskii, S. Rouf, 2021, Dynamics of SIR model with vaccination and heterogeneous behavioral response of individuals modeled by the Preisach operator, J. Math. Biol. 83, 1-34. [CrossRef]
- M. A. Krasnosel’skii, A.V. Pokrovskii, 1989, Systems with Hysteresis, Springer.
- P. Krejčí, K. Kuhnen, 2001, Inverse control of systems with hysteresis and creep, IEE Proceedings-Control Theory and Applications 148 (3), 185-192. [CrossRef]
- K. Kuhnen, 2003, Modeling, identification and compensation of complex hysteretic nonlinearities: A modified Prandtl-Ishlinskii approach, European Journal of Control 9 (4), 407-418. [CrossRef]
- I. D. Mayergoyz, 2003, Mathematical Models of Hysteresis and Their Applications: Electro-magnetism, Academic Press.
- I. D. Mayergoyz, C. Korman, 2019, Hysteresis and Neural Memory, World Scientific.
- F. Preisach, 1935, Über die magnetische Nachwirkung, Z. Phys. 94, 277-302. [CrossRef]
- X. Tan, J. S. Baras, 2004, Modeling and control of hysteresis in magnetostrictive actuators, Automatica 40 (9), 1469-1480. [CrossRef]
- A. Visintin, 1994, Differential Models of Hysteresis, Springer.
- L. M. Werner, 2018, Hysteresis losses in the Preisach framework, Empirical Economics 58, 1249-1278. [CrossRef]
- Subscription Service Statistics and Costs, 2024, C+R Research Report, https://www.crresearch.com/blog/subscription-service-statistics-and-costs/.
- FTC Final Click-to-Cancel Rule, 2024, https://www.ftc.gov/news-events/news/press-releases/2024/10/federal-trade-commission-announces-final-click-cancel-rule-making-it-easier-consumers-end-recurring/.
|
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).