You are on page 1of 14

Journal of Health Economics 47 (2016) 5063

Contents lists available at ScienceDirect

Journal of Health Economics


journal homepage: www.elsevier.com/locate/econbase

Health insurance and diversity of treatment


David Bardey a,b , Bruno Jullien c , Jean-Marie Lozachmeur c,
a
University of Los Andes Cede (Bogota), Colombia
b
Toulouse School of Economics, France
c
Toulouse School of Economics, University of Toulouse 1 Capitole (CNRS), France

a r t i c l e i n f o a b s t r a c t

Article history: We determine the optimal health policy mix when the average utility of patients increases with the supply
Received 26 November 2014 of drugs available in a therapeutic class. Health risk coverage relies on two instruments, copayment and
Received in revised form 9 October 2015 reference pricing, both of which affect the risk associated with health expenses and diversity of treatment.
Accepted 4 January 2016
For a xed supply of drugs, the reference pricing policy aims at minimizing expenses, in which case the
Available online 17 February 2016
equilibrium price of drugs is independent of the copayment rate. However, with an endogenous supply of
drugs, diversity of treatment may susbtitute for insurance so that the reference pricing may depart from
Keywords:
maximal cost-containment in order to promote entry. We next analyze the determinants of the optimal
Health insurance
Drugs market
policy. While an increase in risk aversion, or in the side effect loss, increases diversity and decreases the
Price regulation copayment rate, an increase in entry cost decreases both diversity and the copayment rate.
2016 Published by Elsevier B.V.
JEL classication:
I18
L11
L15
L51

1. Introduction In terms of budgetary concerns, the key question that a health


system faces is not only how much insurance coverage to provide
A controversial issue has long been whether or not costly incre- but also what level of diversity of treatments to offer for a given
mental innovation leading to me-too or follow-on drugs (as pathology. In line with the literature on horizontal differentiation
opposed to major innovation leading to breakthrough products) with Bertrand competition (e.g. see Anderson et al., 1995), one may
has some value for society as a whole. As argued by Wertheimer object that the number of products at equilibrium in a unregulated
et al. (2001), me-too drugs have several advantages, among which market is excessive from a social welfare point of view. This argu-
differing dose delivery systems and dosage forms that enable ment would be reinforced in the pharmaceutical market as entry is
extended uses with a variety of patient population; availability of subsidized via generous insurance coverage. This would ultimately
choice when patient response to and tolerance of a particular agent plead in favour of a strong price control in order to discourage entry
is subject to great variation; [and] the ability to tailor therapy to of me-too drugs (and possibly encourage major innovation). Our
the needs and preferences of patients (pp 78). In line with these contribution is to challenge this view, namely that strong price con-
comments, Di Masi and Paquette (2004) show that among 72 ther- trol should limit the entry of me too drugs. We show that in a
apeutic classes, one third of the me-too drugs received a priority model where individuals are sufciently risk averse with respect to
rating from the US FDA.1 Moreover, 57% of these classes include a poor health conditions, entry of me too drugs should be encour-
me-too drug that received such a priority rating. As such, diver- aged via soft price controls even when there is an (optimal or non
sity of treatment itself has an insurance role in that it increases the optimal) insurance plan in place.
probability that a patient nds the best (or the least bad) treatment. We analyze this issue in a context where health services are
supplied in an imperfectly competitive market, as is the case for
drugs markets. In order to analyze the issue of the diversity of
treatments offered to policyholders, we focus on two policy instru-
Corresponding author. Tel.: +33 5 61 12 87 37. ments that are used to maximize individuals expected utility. The
E-mail address: jean-marie.lozachmeur@tse-fr.eu (J.-M. Lozachmeur). rst is a standard linear copayment rate, while the second is a ther-
1
Drugs receiving priority review (as opposed to standard review) by FDA benet
from an accelerated procedure for approval (6 months as compared to 10 months
apeutic reference pricing regulation. We use the latter instrument
under standard review). as it provides a simple characterization of the degree of the price

http://dx.doi.org/10.1016/j.jhealeco.2016.01.003
0167-6296/ 2016 Published by Elsevier B.V.
D. Bardey et al. / Journal of Health Economics 47 (2016) 5063 51

control. Moreover, therapeutic reference pricing is used more and weight attached to the lowest price leads to a reduction in prices
more frequently in developed countries2 and its effectiveness has and increases the market share of the cheapest drug (the generic
been demonstrated (e.g. see Brekke et al., 2009, 2011). While terms drug in their context).
vary across countries, internal or therapeutic reference pricing (as In a model that distinguishes breakthrough and follow-on drugs,
opposed to external reference pricing) consists of determining a Bardey et al. (2010) evaluate the long run impact of reference pri-
reference price as a weighted sum of drugs prices adopted in the cing on pharmaceutical innovation, delays of introduction, patients
same therapeutic class. If the price of a drug is higher than this ref- health and expenditures. They show that reference pricing reg-
erence price, patients pay the full difference between the price of ulation yields lower prices and therefore delays pioneer drugs
the drug and the reference price. This regulatory scheme can be and me-too entries. Nevertheless, as me-too entries are more
perceived as a complement to copayment rates in order to encour- delayed, it may favor costly breakthroughs and may increase health
age patients to consume low price drugs. expenditures in the long run. Miraldo (2007) compares the equi-
We build a model where there are several pharmaceutical rms librium allocations under various alternative regimes: copayment
selling horizontally differentiated drugs that belong to the same and reference pricing schemes.6 In contrast, our model is more nar-
therapeutic class.3 All patients value the drugs differently because row on the reference pricing modality as we focus our attention on
of their different side effects.4 Risk averse consumers benet from internal reference pricing, however the optimal copayment rate
an insurance plan consisting of a premium and a linear copayment and the drugs pricing regulations are jointly determined.
rate subject to internal reference pricing. We consider a reference On the normative side, very few papers study health insurance
price that is a linear combination of extreme value prices in the in the context of pharmaceutical markets. Lakdawalla and Sood
market. By choosing the weight attached to the lowest price in (2009) show that encouraging health insurance can be welfare
the reference pricing formula, the regulator determines the pres- improving as it lowers static deadweight loss (i.e. it implies more
sure on the equilibrium price of drugs and thus on total health efcient utilization) without altering incentives for innovation.
expenses.5 Our results show that in the short run, the reference They also show in another paper (2013) that a competitive health
pricing scheme aims at minimizing the price of drugs. Indeed, as insurance market can combine static and dynamic efciency in the
long as the number of drugs in the therapeutic class is xed, the drugs market: a premium-nanced xed fee is offered to drugs
health insurer is only willing to lower drugs prices since he can- monopolists which ensures second best utilization and extracts
not improve the drugs diversity. However, in the long run, there the full surplus so that incentives for innovation are optimal. Their
may be room for a more lenient reference pricing policy accommo- model however, does not provide for the optimal diversity of
dating some increase in health expenses in order to improve the treatment since individuals can only benet from at most one inno-
diversity of treatments. The desirable level of the reference price vation.
is the result of a trade-off between a diversity effect (where a new The next section presents the set-up. Section 3 is devoted to
drug decreases average side effects) and the xed cost generated the short-term analysis assuming a xed number of drugs. Sec-
by additional drugs entries. We particularly emphasize the role of tion 4 characterizes the optimal regulation in the long run, with an
policyholders risk aversion by showing that the higher the level of endogenous number of drugs. Section 5 discusses some possible
risk aversion, the more likely it is that the regulator chooses a more extensions. Lastly, Section 6 concludes.
lenient reference pricing policy. While an increase in risk aversion
leads to a lower copayment rate and a higher diversity of treat-
ments, an increase in innovation costs implies a lower copayment 2. The set-up
and a lower diversity of treatments.
Consider a collection of policyholders J (0, 1) who can be
1.1. Related literature unwell with probability  and healthy with probability 1 . We
normalize the size of the population to 1. In case of illness, each
While mainly normative, our article borrows from the positive patient chooses a drug among N ( 2) treatments, denoted by a
literature dealing with drugs price regulation. To the best of our subscript i I (1, N). We refer to N as the diversity of treatment.
knowledge, this is the rst paper that endogenizes both the ref- When choosing a drug i in case of illness, a patients net income is
erence pricing regulation and insurance scheme in a fully-edged wsi = w  pi ,7 where w is an exogenous income,  denotes the
equilibrium model. Brekke et al. (2007) develop a general set-up premium paid to the health insurer and pi is the out-of-pocket price
containing horizontal and vertical differentiations in drugs mar- paid for consuming drug i. When healthy, the net income is wh =
kets. They consider two reference pricing rules, namely internal and w . In the state of illness, policyholders are horizontally differ-
external reference pricing as well as price cap regulation, with both entiated: when consuming drug i, a policyholder j J is affected by
regulatory schemes being associated with an exogenous copay- a side effect which depends on an individuals stochastic variable
ment rate. Concerning the internal reference pricing, they consider xji X, observed by the policyholders before consumption takes
the laboratories best reply function, and provide comparative sta- place. The shocks xji are identically and independently distributed
tics on the equilibrium allocation. They show that an increase in the across policyholders and drugs over R+, and follow an exponential
distribution with distribution function F (x) = 1 exp (x) . This
assumption implies that the mass of individuals with large adverse
2
See Lopez-Casanovas and Puig-Junoy (2000), Danzon (2001) and Danzon and effects is always lower than the mass of individuals with small
Ketchman (2004) for more details on reference pricing and its applications.
3
adverse effects.8
This horizontal differentiation set-up can capture two types of competition in
drugs market. It accomodates competition between brand-name drugs, i.e. pioneer
drug and me-too drugs, which belong to the same therapeutic class. It also ts the
competition between generic drugs which reproduce the same molecule (horizontal
6
differentiation occurs because formulation, packaging as well as delivery systems A recent work by Brekke et al. (2015a) analyses a similar issue in a set-up
are allowed to vary). Conversely, an horizontal differentiation framework is not well encompassing vertical and horizontal differentiation. The authors test the empirical
adapted to model competition between brand-name and generic drugs. validation of their model in a companion paper (Brekke et al., 2015b).
4 7
Side effects accommodate secondary effects per se, adverse drug reactions, drug- There is no lump sum monetary compensation for being sick. This may be ruled
drug interactions, dosing schedules as well as delivery systems. out by law or because the state of illness is not veriable by the regulator.
5
In some European markets, such as Germany, the weight also depends on supply 8
Our main result can be shown to hold if the distribution is such that 1 F (x) is
conditions and market shares. We thank a referee for pointing that to us. log-log concave (see Bardey et al., 2013)
52 D. Bardey et al. / Journal of Health Economics 47 (2016) 5063

The policyholder js utility when unwell and consuming drug i where


is:
1
  Q =    < 1.
Uji = u wsi  xji , E exp mini I xji

where  > 0 is a xed effect of illness on the VNM utility function


when unwell.9  measures the side effects marginal loss (which The parameter Q is a measure of the average side effects of available
corresponds to the differentiation parameter in traditional models treatments which impacts the average utility when unwell. Given
with horizontal differentiation) so that  xji represents the mon- that u is negative (due to the CARA specication), the expected util-
ity increases when Q increases. To ensure niteness of the expected
etary equivalent loss incurred by a patient j consuming drug i.
utility for all N we state the following:12
There is also an outside alternative which is to forgo any treat-
ment: 0 xj0 denotes the monetary equivalent loss incurred by an Assumption 1. The market is such that N >  .
unwell individual when not consuming any drug. xj0 is assumed Since x = minxji is exponentially distributed with mean 1/N,
iI
to be exponentially distributed along with the xji . Throughout the Assumption 1 ensures that:
paper (except in Section 5.1), we consider that  0 is innite so that

the market is fully covered. Roughly speaking, the cost of not being 1


= exp x N exp [Nx] dx,
treated is so large compared to prices and adverse effects that all Q (N) 0
policyholders buy a drug when unwell.10 Throughout the paper, we
use an exponential VNM utility function of the CARA form: is well dened so that:

u () = exp () , 
Q (N) = 1 . (3)
N
where  represents the absolute risk aversion parameter. This spec-
ication allows us to avoid any income effect on the demand for Our model relates the average side effects to the diversity of avail-
drugs which simplies the exposition of our results and allows us able drugs in the therapeutic class. We thus refer Q (N) as the
to perform simple comparative statics with respect to the risk aver- welfare index of diversity (WID hereafter). Ceteris paribus, the higher
sion parameter. We show in Section 5.2 that our main results hold the number of drugs available in the therapeutic class, the higher
when the adverse effect enters separately in the utility function. the perceived utility of the health service as patients benets from
Throughout the paper, we consider a sequential game where the a decrease in the average side effects. This marginal benet from
regulator moves rst and sets a reimbursement policy consisting a a reduction of side-effects decreases in the number of drugs. On
copayment rate and a reference pricing scheme (described in the the contrary, the negative consequences of side effects on patients
next section). The rms decide to enter or not the market and then welfare increase both in the side effects marginal loss  and in the
compete in prices. To sum up, the precise timing of the game is as risk aversion parameter .
follows:

3. Insurance policy and the short-run equilibrium of the


1. The regulator chooses the reimbursement policy (copayment a pharmaceutical market
and reference pricing r dened in the section) and the insurance
premium . In this section, we rst determine the market equilibrium in
2. Drug producers enter or not. the therapeutic class and then analyze the short-run insurance
3. The N active drug producers simultaneously set prices. trade-off when diversity is given (there is no entry at stage 2). We
4. A fraction  of consumers become ill. Each unwell consumer endogenize the diversity of treatment in the next section.
privately observes the vector of individual effects xji , and decides
which drug to buy (or not to buy).11
3.1. Insurance and market equilibrium

At a symmetric equilibrium where pi = p for every i (1, N), We consider one specic therapeutic class in which pharma-
patients minimize the loss so that the aggregate expected utility ceutical rms each own one drug and compete in the market for
function writes: drugs by setting prices. While prices are freely chosen by rms,
   the equilibrium is shaped by the insurance policy which is based
EU = E u ws minxji + (1 )u(wh ), (1) on two instruments: a copayment rate a [0, 1] and reference
iI
pricing.13 We model reference pricing as a rule that associates
where E is the expectation operator over J XN . any vector of manufacturers prices (P1 , . . . PN ) with a reference
For a CARA utility function, Eq. (1) yields: price R (P1 , ..., PN). A patient buying drug i is thus reimbursed
(1 a) min Pi , R , that is a share 1 a of his expense up to R,
 but obtains no reimbursement for the excess payment above the
EU =  u (ws ) + (1 )u(wh ), (2)
Q reference price.
Several formulas for the reference price can be used. We con-
sider an internal reference pricing regulation by assuming that the
9 reference price is a weighted sum of the minimal and the maximal
The value of  can be lower than one as suggested by Finkelstein et al. (2013). As
we show later, this parameter only plays a role in the detemination of the insurance
coverage.
10
This assumption is quite realistic in countries characterized by high levels of
12
coverage, i.e. low levels of out-of-pocket prices. In the long-run analysis, the assumption constrains an endogenous variable N.
11
An equivalent assumption is to assume that the doctor perfectly observes the However, Assumption 2 below ensures that it holds in any equilibrium.
13
state of the patient and delivers good information on possible adverse effects of any Alternative mechanisms to reference pricing could be price-cap or direct price
drug to ensure the best match between each drug and each patient. negotiation with producers.
D. Bardey et al. / Journal of Health Economics 47 (2016) 5063 53

prices of the drugs in the market.14 Formally, the reference price is Proposition 1. For any N and policy (a, r), there exists a unique
given by: symmetric price equilibrium with a drug price
  (N)
R = rmin Pi + (1 r)max Pi , (4) P= ,
i i
where r [0, 1] is the policy variable, referred to as the strength of where (N) = / (N 1) is the inverse semi-elasticity
  residual
of the
the drug price regulation. When r = 1, there is maximal reference demand evaluated at pi = p(i.e. D (0, N) / D (0, N) /pi ).
pricing. This is the most stringent policy from the perspective of
pharmaceutical rms. Conversely, when r = 0, there is no reference Proof. See Appendix A.
pricing. In such a case, all the treatments are reimbursed at a rate The resulting price equilibrium is identical to the one in a
1 a. standard general model of horizontal differentiation in which the
We rst discuss the last competition stage of the model (stage 3 differentiation parameter would be /. Here, the pass-through rate
and 4 mentioned above). The number of drugs N is given as well as increases in the strength of reference pricing and in the copay-
policy parameters (a, r, ). As already mentioned, in the main body ment rate. In other words, increasing the insurance coverage lowers
of the paper, we assume that the market is fully covered and relax the competition intensity. This feature of the model illustrates what
this assumption in Section 5.1. The demand for one drug is thus the Danzon (2012) refers to as supplier moral hazard. Equivalently,
mass of individuals who prefer this drug to any other. We focus on relaxing the strength of reference price regulation lowers the com-
symmetric equilibria where rms set the same prices. petition intensity.
Consider any drug i (1, N) and suppose that the net out- At a symmetric equilibrium, the pharmaceutical rms prot is
of-pocket price paid by patients for all other drugs is p. Our then:
assumptions on the utility functions (in particular, the fact that the
utility of unwell patients are symmetric and quasi-linear in prices)  (N)

(N) = . (5)
imply that the demand Di , when the out-of-pocket price paid for N
consuming drug i is pi , can be expressed as follows at any symmetric The closed form solution is useful in order to show in a more obvious
equilibrium: way the different effects at work. Ceteris paribus, the side effects
marginal loss  increases the drugs prices and the prot obtained by
Di = D (pi p, N) , laboratories. However, the overall differentiation (N) decreases in
the diversity of treatments available in the therapeutic class.
where D is a function independent of i with D/pi < 0 for any pi ,
p and N. At a symmetric equilibrium, each pharmaceutical rms
market share is D (0, N) = /N.15 3.2. Fixed drugs diversity
By denition, when all other rms charge P and rm i charges
Pi , the out-of-pocket price paid by patients is dened by: Here we determine the optimal policy in the short-run, consid-
ering the market structure as given. Since the diversity N of
pi = Pi (1 a) min [Pi , R] available drugs in the therapeutic class is xed, we drop the argu-
ment N in the functions of this section. The resource constraint
so that, of the regulator states that the premium should cover reimburse-
ment cost that is  = (1 a)(1 + )P where is to be interpreted as
pi p = Pi P (1 a) (min [Pi , R] min [P, R]) .
a loading factor or as a shadow cost of public funds. Thereafter, we
Using the denition of R in Eq. (4), this yields: consider two alternative scenarios. First, we study the case where
the social objective is to maximize the policyholders expected util-
pi p = (Pi P), ity. Second, as a robustness check, we integrate industry prots in
the regulators objective.
where = a + (1 a)r measures the distortion introduced by the
insurance system on the market for drugs. We refer to as the
3.2.1. Social objective without laboratory prots
pass-through rate resulting from the joint regulation of insurance
When the social objective is to maximize the policyholders
coverage and reference pricing. This rate determines the fraction
expected utility as dened in (2).16 For a xed diversity of treat-
of the price differential which is actually paid by individuals and is
ments, the optimal policy is the solution of:
increasing in the copayment rate and the reference pricing strength.
Note that our reference pricing specication does not generate any 
max EU = (1 )u(w ) +  u (w  aP)
kink in the demand function, as opposed to Danzon (2001). This is 0a,r1 Q
due to the fact that we focus on symmetric equilibria in fully cov-
ered markets. In Section 5.1, which considers non-fully covered s.t  = (1 + )(1 a)P, P= .
a + (1 a)r
markets, such a kink emerges.
For-prot pharmaceutical rms compete in prices, and marginal For any r < 1, a decrease in a exacerbates the extent of supplier moral
costs are normalized to 0. Therefore, we have: hazard by increasing the price paid by patients. Nevertheless, this
problem can be dealt with by adjusting the strength of reference
Pi arg max Pi D ( (Pi P) , N) . pricing. For a xed supply of drugs, reducing the price P can only
benet consumers because the WID is xed and the price is always
As the symmetric equilibrium yields Di = /N, we obtain:
above marginal cost. As increasing the weight r in the reference
price decreases the equilibrium price, it follows that it is always
optimal to choose maximal reference pricing r = 1. The price is then
14
Even though each country may have some specicities regarding its applica- at , the level that would prevail without any insurance scheme. In
tion, 20 of the 27 EU Member States had implemented an internal reference price
 and
regulation (see Ruggeri  Nolte, 2013).
15
More formally 1/ D (pi p, N) is the probability that the total loss  xji +
16
pi be less than minxjl + p. This alllows us to compare our results with the ones in the following section
l=
/ i where prots will be nil.
54 D. Bardey et al. / Journal of Health Economics 47 (2016) 5063

and 1, so that there is no need for insurance for (1 ) / and


lambda 2.5 there is full insurance for 0 .
2.0
no insurance 3.2.2. Social objective with laboratory prots
1.5 partial Let us now assume that the policyholder holds a fraction [0,
1] of laboratories prots. Since prots are equal to P, the policy-
1.0
holders expected utility becomes:
0.5 full insurance 
EU = (1 )u(w  + P) +  u (w  aP + P) .
Q
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 The only change with respect to the previous section is the change
WID
of net income due to prot redistribution (the equilibrium demand
Fig. 1. Optimal insurance scheme with xed supply. for drugs is unchanged so that the price P and the premium  remain
the same as well as the WID). We now show that maximal reference
other words, full reference pricing allows the supplier moral hazard pricing still applies. To see this, note that r only matters when there
effect to be cancelled out. is some insurance, i . e . when a < 1 . Then, using  = (1 a)(1 + )P,
Given that the price is xed at P = , the problem is reduced to we can write the net income in the two states as:
a standard insurance problem. The slope of the objective function wh = w (1 + )P + (1 + )aP,
with respect to a is:
ws = w (1 + )P + ((1 + ) 1) aP.
EU 
= EU (1 + )P u (ws ) P, (6) Now suppose that r < 1 and a < 1 . For given level of aP, the regula-
a Q
  tor can reduce P by increasing r and raise a. This change increases
where EU = (1 ) u (wh ) + /Q u (ws ) is the policyholders the net income in both states of nature and thus unambiguously
expected marginal utility of income. Rearranging (6) for a solution increases social welfare. It follows that at the optimal policy, either
with positive insurance coverage gives: r = 1 or a = 1 (in which case r = 1 is also optimal). Given that the
u (ws ) optimal strength of reference pricing is r = 1, the price is at its com-
EU (= 0 if a > 0). (7) petitive level P = . Thus the choice of copayment does not affect the
Q (1 + )
price P but only the level of insurance. As there is no wealth effect
The optimal policy is then characterized in the following proposi- in this model, redistribution of prots  amounts to change the
tion: wealth w to w +  that has no effect on the optimal copayment
Proposition 2. For a xed diversity of treatments N, the policy. More formally, we have
optimal strength of reference pricing is r = 1 and the copay-
EU EU
ment rate is non-decreasing in the loading factor . If 0 = | >0 = u ( ) | =0 .
a a
( 1) (1 ) / (1  + ), full insurance is optimal for all N.17 If
> 0 , there exists N such that no insurance is optimal if diversity N We thus conclude that the optimal policy is the same as with no
is such that N N, while the optimal copayment rate a is implicitly prot redistribution. While our conclusion for reference pricing
determined by (7) if N < N. is very robust, the conclusion for copayment relies on the utility
function.
Proof. See Appendix B.

For a small loading factor , the increase of the premium is not 4. Endogenous drugs diversity
sufcient to require a positive copayment rate. This is due to the fact
that the side effects render the marginal utility of income depend- In this section, we endogenize the market structure in the thera-
ent on the state, and in particular the marginal utility larger when peutic class, and we focus on the free-entry equilibrium that occurs
unwell.18 For larger than 0 , the risk when N is large is not suf- in the long run (stage 2 of the timing presented in Section 2).
cient to grant insurance. Notice that when N is large, the price is Laboratories enter into the market as long as the prots earned
close to zero and the average side effect is also close to zero, but in the drug market outweigh a uniform sunk cost k generated by
there may be some residual risk due to the direct consequence of research and development activities.21 For conciseness we ignore
illness on utility (as measured by ). Thus when  is small, the opti- the fact that N is an integer and treat N as a continuous variable (see
mal policy is to provide no insurance at all. Partial insurance occurs Footnote 23 for more details).
when the supply of drugs is insufcient and in this case, the opti- Equalizing the sunk cost to equilibrium prots given by Eq. (5),
mal copayment rate, determined by condition (7), is the result of a the diversity for a free-entry equilibrium is implicitly given by (up
trade-off between risk pooling (ex-ante efciency) and an increase to the integer):
in the size of the risk (ex-post efciency).19
 (N)
The proposition is illustrated in Fig. 1 which represents the k= . (8)
N
choice in the space (Q (N) , ).20 There are two curves such that
full insurance is optimal below the lower curve while no insurance Accounting for endogeneity of supply changes the nature of the
is optimal above the upper one. The two curves coincide at zero analysis compared to the case with xed supply. Indeed, when N

17 21
Of course, if  < 1, this cases never applies. This assumption ts well with reality, as incremental innovation has often been
18
In this case a policy that would provide over-insurance would be optimal but generated by unintentional research aimed towards producing breakthrough prod-
this option is ruled out. ucts (see Wertheimer et al., 2001). One can however argue that, starting from a
19
See Geoffard (2006) for a complete analysis dealing with this trade-off in a certain number of drugs existing in a therapeutic class, the xed cost of innovation is
context where ex-post moral hazard is due to a quantity effect. decreasing since me-too drugs enter the market with less clinical trial experiments
20
It is drawn for  = 1,  = 1.5,  = 1/3 and = 1.2 . Thus the diversity index is Q = 0.5 (see Frothingham, 2004). A decreasing cost function would ultimately reinforce our
for N = 3. result stating that less than maximal reference pricing would be optimal.
D. Bardey et al. / Journal of Health Economics 47 (2016) 5063 55

is xed, the only scope for price policy is cost containment. Endo- By denition of N, we have that for a < 1 : kN =  (N) > akN.
genizing the supply then creates a linkage between total expenses Therefore the set of admissible N is a non-empty compact inter-
and diversity. The free-entry condition implies an increasing rela- val so that an optimal diversity exists. Denote by ra the strength
tionship between diversity of treatment and expenses: of reference pricing as a function of a. When r = 1, the level of entry
is always equal to N. Hence, at any copayment level, it is possible
P = kN. to implement the minimal WID at the minimal cost. One may raise
While the health policy may raise diversity, it has to account for diversity above this level by reducing the strength r of reference pri-
this cost. This suggests that we view the choice of health policy as cing. However, this would imply some more entry costs. The rst
the joint choice of insurance, and diversity of treatment with a cost question is thus whether or not it is optimal to minimize cost and to
function that is determined by the free-entry condition. rely only on insurance, or to accommodate some cost increase. The
Given that = a + (1 a)r 1, the minimal number of labora- gain of raising diversity is captured by the slope of EU with respect
tories entering the market is N given implicitly by the free entry to N which is equal to:
condition (8) with = 1 or: EU  Q 
= 2 u(ws ) k (1 + ) (1 a) EU aku (ws ) ,
 N Q N Q
N (N 1)
, (9)
k where EU is dened as in Eq. (6). Using Eq. (7) and u = u, this
where we used (N) = / (N 1). Recall that Q is dened only for slope can be rewritten as:
N > . To ensure this condition we assume (dening N = + if     EU
0 ): EU  Q  1a
= k u (ws ) |N , (11)
N Q N Q N a
Assumption 2. N 2 and  < N < N.
where EU/a|N is given by the LHS of Eq. (6). The rst term in
Assumption 2 guarantees that there are at least two rms in the RHS of Eq. (11) captures the trade-off between a higher WID
the market and the WID is well dened (Q > 0). Moreover, the (associated with more entry) and the cost of entry. The last term
condition N < N ensures that there will be some insurance in captures the extent to which reference pricing can correct for insuf-
equilibrium, i.e. the copayment will be less than 1.22 cient or excessive copayment to provide some insurance. Indeed,
In order to provide more intuitions about optimal reference pri- when there is insufcient coverage (i.e. EU/a|N < 0), additional
cing, we proceed in two steps. First, for any given copayment rate entry acts as a substitute to insurance since it reduces the difference
a, we determine the optimal strength of reference pricing. In a sec- of marginal utilities between healthy and unwell individuals. Thus,
ond step, we determine the joint optimal choice of copayment and entry is more valuable when the copayment is too large. Whether
reference pricing. it is optimal to reduce the strength of reference pricing to induce
some entry depends on the level of the copayment rate.
4.1. Reference pricing and diversity for xed copayment
Proposition 3. Suppose the copayment rate a is xed.
Even though we consider this step in order to more easily decon- If  1, the reference pricing strength ra < 1 and the optimal
struct the different effects at work, it is worth stressing that such a diversity N is above N for a (a, 1) where a < 1. Moreover ra < 1
situation may occur when the copayment rate is set by another reg- for all a if full insurance is optimal in the short-run at N = N.
ulator, or at a high level of aggregation of pathologies/therapeutic If  < 1, either diversity N is minimal for all a (i.e. ra = 1), or ra < 1
class. For a given copayment rate, the strength of reference pricing for a (a, a) with 0 a < a < 1.
determines the equilibrium price for any level of entry. Increasing r
Proof. See Appendix C.
results in lower equilibrium prices and consequently in less entry.
Hence, the optimal policy reects a trade-off between the total cost When risk aversion and the side effects marginal losses are not
of treatment and the WID. too small, it is optimal to induce entry by relaxing the intensity of
Notice that the free-entry equilibrium (8) induces a decreasing competition at least for large levels of coverage. This conclusion
relationship between r and N. Hence, for a given a, choosing r is contrasts with the welfare analysis of entry in standard models
equivalent to choosing the diversity N under the relations: of horizontal differentiation, which conclude that there is always
excessive diversity at an unregulated free-entry equilibrium (e.g.
 (N) /kN a
0r= 1 and P = kN. see Anderson et al., 1992). The key difference between our setup
1a
and these models is that diversity has an additional benet in a
Thus, the problem reduces to the choice of diversity within a fea- pharmaceutical sector: improving the welfare of unwell individ-
sible range, and the optimal policy is the solution of the following uals improves the overall insurance performance. This is the case
program:23 in particular when full insurance is optimal, as alternative means

 kN
 of providing insurance helps to reduce the gap between marginal
maxEU = (1 )u (wh ) + u wh a , utilities. On the other hand, for low levels of differentiation and risk
N Q 
aversion, the insurance benet is not so important and the tradi-
s.t. wh = w (1 + )(1 a)kN, tional result of excessive entry may apply. In this case, relaxing the
(10)
strength of reference pricing may be desirable only for intermediate
N N, levels of coverage, or not at all.
akN  (N) 0. We show in Appendix D that the objective is quasi-concave
which allows us to perform simple comparative statics.

Proposition 4. For a xed copayment a, the optimal diversity N:


22
To see this, note that N = N if a = 1. The condition implies that a < 1 is optimal
for N = N, and therefore it is impossible that a = 1 is an optimal choice. i) is non-decreasing in the copayment rate a if there is excessive cov-
23
When N is treated as a discrete variable, several values of r generate the same erage, i.e. if EU/a 0;
N . Clearly it is optimal to choose the maximal strength compatible with N, hence r
is such that P = kN/ .The problem is thus identical but restricted to integer values
ii) is non-increasing in the loading factor and non-decreasing in the
of N . The comparative statics are the same. risk aversion  and the adverse effect risk .
56 D. Bardey et al. / Journal of Health Economics 47 (2016) 5063

Proof. See Appendix D. Proof. See Appendix E.

Increasing the copayment has two effects. It reduces the level Increasing always leads to less coverage and less diversity as
of insurance and the extent of the suppliers moral hazard. Since both are costly. Still, whenever  is large enough, it is optimal to
entry induces some indirect form of insurance through a higher raise diversity above the minimum by relaxing the strength of refer-
WID, the two effects make entry more attractive if there is not ence pricing and inducing some entry. Conversely, if  is smaller
enough insurance. In such a case, the optimal diversity of treat- than 1, cost minimization tends to prevail and minimal diversity
ment increases in the copayment. However, if there is excessive will be chosen for not too small.
insurance, the two effects work in opposite directions and no con- As the copayment rate is non-decreasing in , the proposition
clusion can be derived. The other results outlined in Proposition 4 also reveals that there exists a threshold such that there is par-
are rather more intuitive, and state that the policy becomes more tial insurance if > , and in this case the policy departs from
oriented toward cost containment when the cost of public fund cost minimization when the combined value of risk aversion and
increases, risk aversion decreases or the differentiation parameter disutility from side effects is large enough.
decreases. When < , the optimal policy coincides with the policy ana-
lyzed in the previous section for a = 0. On the range > , we
4.2. Optimal policy mix can analyze how the optimal level of copayment and the optimal
diversity of treatment offered in the therapeutic class varies with
We now consider the complete set of instruments and discuss preferences and the xed cost.
the optimal choice of coverage and strength of reference pricing,
Proposition 6. For > (a > 0) and  > 1 :
(a , r ). At the optimum, r = ra which exists for all a . The value of
the program (10) is continuous in a so that an optimal copayment a*
i) An increase in the risk aversion  or the adverse effect risk 
exists. As already mentioned, Assumption 2 rules out the possibility
increases diversity N and decreases the copayment rate a;
of no insurance. Hence, we only need to consider full insurance or
ii) An increase in the xed cost k decreases both diversity N and the
partial insurance. Before determining which regime prevails, we
copayment a;
discuss the strength of reference pricing in each regime.
iii) An increase in the marginal utility of income when unwell 
Proposition 3 implies that for  larger than 1, a full insurance
decreases the copayment rate a but does not affect diversity N.
policy is necessarily associated with some incentive to entry and
r* < 1 . To see this, note that if a* = 0 and r* = 1, then N = N. But this
Proof. See Appendix F.
would contradict Proposition 3 which implies that r0 < 1 (as full
insurance must be optimal for a* = 0).  The results
 rst state that more risk aversion () or more risk
Let us now consider the case where partial insurance is optimal.  and  imply more coverage. This is in line with standard insur-
The rst-order conditions with respect to N give (using Eqs. (6) and ance analysis. The result on k follows the same logic: as providing a
(11) for EU/a = 0): given level of WID becomes more costly, the treatment expense
increases and the optimal policy compensates for the risk by a
u (ws )
EU = ; (12) reduction in the copayment rate. The result i) shows that insurance
Q (1 + )
and the diversity of treatment are complementary instruments to
 Q accommodate an increase in risk aversion or in the heterogeneity of
k (= if N > N). (13) side-effects. However, result ii) states that they evolve as a substi-
Q N
tute when the cost of entry increases. The last result is interesting as
Eq. (12) describes the same trade-off between risk pooling and size
it states that the optimal diversity is independent of . What mat-
of the risk as in the benchmark section, but obviously taken at a
ters is the marginal increase in utility that diversity of treatment
different value of N. The second equation captures the trade-off
induces, and not the absolute value of disutility.
between the social benets and costs generated by the entry of a
new drug, respectively described by the LHS and the RHS of Eq. (13).
4.3. Numerical illustration
The social marginal cost is simply the welfare effect of an additional
sunk cost, accounting for the optimal allocation of this cost across
In order to illustrate the results of this section we run different
health states. The social marginal benet captures the increase in
simulations. Our goal is to illustrate the trade-off between costs
the WID through the marginal reduction of average adverse effects
and social benets from an additional entry of a drug. Remember
generated by a new drug entry. To follow, we analyze whether or
that the social cost is the xed cost k while the social benet comes
not the optimal diversity of treatment is interior, i.e. whether or not
from the increase in the WID. We thus run simulations for different
there is less than maximal reference pricing.
values of k, and for each value of k we run three simulations depend-
Recall that the minimal diversity N is given implicitly by the
ing upon the value taken by , the latter being positively related to
free-entry condition k =  (N) and is independent of the level
the WID. The other parameters are xed to  = 0.8,  = 0.5, = 0.3,
of absolute risk aversion . Substituting (13) in (12), a necessary
w = 1 and  = 0.05. For each scenario, we report the long run val-
condition for an interior solution for r is:
ues of n* , a* and r* of Section 4.2. We also provide the value a above
 Q (N) which it is optimal to introduce one additional entry as described
k > 0. (14)
Q (N) N in Proposition 3. Recall that it is optimal to induce an additional
entry when the exogenous level of copayment is above a.
We can now state our main proposition:
 Table 1 presents the results. In a rst scenario, we have k = k1 =
Proposition 5. The optimal copayment rate a* is non-decreasing in   /2. As described in (9), this is the highest possible xed cost
the loading factor and diversity N* is non-increasing in . that implies a minimum number of rms exactly equal to N = 2
when reference pricing is at its maximum  strength.
 The value of
i) If  1, then the reference pricing strength r* < 1 and diversity is k is then successively decreased to k2 =   /2.5 and to k3 =
 
above N for all ;   /4. These latter values still imply a minimum number of
ii) If  < 1, then the reference pricing strength is maximal, i.e. r* = 1 rms equal to 2 (the maximum value  of kimplying a minimum
when the copayment rate a* > 0. number of rms equal to 3 is equal to   /6). For each of these
D. Bardey et al. / Journal of Health Economics 47 (2016) 5063 57

Table 1
Examples with different values of k and .
     
k k1 =   /2 k2 =   /2.5 k3 =   /4

 1.4 1.8 1.9 1.4 1.8 1.9 1.4 1.8 1.9


N* 2 2 3 2 3 3 3 3 3
a* 0.13 0 0.12 0.17 0.18 0.14 0.60 0.29 0.23
r* 1 1 0.12 1 0.28 0.31 0.16 0.53 0.56
a 0.44 0.10 0 0.55 0 0 0.08 0 0

scenario, we report the values of interest variables corresponding in the extent to which this possibility implies more or less diversity
to three different values of risk aversion as measured by . In all at the optimum.
the scenarios, we have  > 1. Dene 0 = / 0 as the measure of the opportunity benet of
By Proposition 5, entry is thus desirable at the margin. However forgoing a treatment. The symmetric equilibrium policyholder js
N is discrete so that an additional entry may not be desirable for expected utility now becomes:
 low enough. When k is at its maximum value compatible with a    
minimum number of drugs equal to 2 (k = k1 ), an additional entry  i
EU = E u wh min minxji + p; x + (1 )u(wh ),
is desirable when  1.9. Below this level, the xed cost is too high iI 0 0
to induce one additional entry. As the xed cost decreases from
where the agent now has access to one more option.
k2 to k3 however, entry becomes desirable for lower values of risk
Suppose that all agents face an out-of-pocket price p except for
aversion.
drug i I. The demand function for drug i is given by the mass of
In all the scenarios, the long-run copayment rate a* is always
unwell individuals who prefer drug i over the others and whose
lower than 1. When an additional entry is desirable (N = 3), a*
severity of illness justies a treatment:
decreases with risk aversion as shown in Proposition 6. How-
 p p N1   p 
ever, when the desirable number of drugs jumps from 2 to 3, the i i
Di =  f (x) 1 F +x 1F 0 + 0 x dx.
copayment rate increases. This is due to the effect of entry of equi-  
0
librium prices. As N jumps from 2 to 3, the price (and the premium)
decreases so that insurance motives becomes less important. We then have:
 
Consider now the reference pricing mechanism implementing  (N 1) (pi p) + 0 pi
Di = exp .
the long-run solution. For each values of k, the value of r jumps N + 0 
from 1 to lower values when the desirable number of rms passes
Using the denition of the reference price mechanism and = a +
from 2 to 3. For example, when k = k2 , the value of r changes from
r (1 a), one thus has:
1 (the most stringent policy) to a value of approximately 1/3 of
 
the maximal price when  increases from 1.4 to 1.9. Alternatively,  (N 1) (Pi P) + 0 (Pi (1 a) min[Pi , R])
Di = exp ,
for a given value of risk aversion of 1.9, the reference price policy N + 0 
becomes less stringent as the xed cost decreases. It increases from
where R is given by (4). The demand function is thus dened by:
12% to 56% of the maximal price when k increases from k1 to k3  
since this is because it is easier to induce entry as k decreases. Now  (N 1) (Pi P) + 0 aP i
Di (Pi , P) = exp if Pi P;
consider the case where k = k3 in which an additional entry is always N + 0 
desirable. One can see that the value of r increases from 16% to  
 (N 1 + 0 ) (Pi P) + 0 aP
56% as  increases from 1.4 to 1.9. This is due to the fact that the Di (Pi , P) = exp if Pi P.
N + 0 
copayment rate decreases with the risk aversion parameter. For a
given value of r, it means that prices and thus prots are higher as As expected, with an elastic aggregate demand for drugs, the refer-
 increases. The value of r implementing a number of rms equal ence pricing system with r > 0 yields a kink in the residual demand
to 3 thus becomes lower. that a pharmaceutical rm faces, as pointed out in Danzon (2001).
Let us nally comment the value of a and remember that This kink implies that the demand is less elastic for a price Pi below
this variable reects the threshold value above which one addi- the market price than for a price Pi above. As a result there exist
tional entry is desirable when a is exogenously set. As shown by multiple equilibria.
Proposition 3, entry is always desirable only when a > a. In par- Proposition 7. Suppose that the opportunity benet of forgoing a
ticular, entry is always desirable when a is exogenously set to 0, i.e. treatment 0 >0. Then
the case when there is full insurance. It is worth to notice that the
 there exists a continuum of symmetric price
equilibria P P, P where
value a is non increasing with the risk aversion for each level of k.
 
P= and P =
(N 1 + 0 ) (N 1) + 0 a
5. Extensions
with = a + (1 a) r.
In this section, we explore two alternative issues related to our Proof. See Appendix G.
set-up. First, we develop the model when the drug market may not
be fully covered. Second, we check whether or not our main results We can measure the size of the equilibrium set by the following
hold with preferences in which side effects enter separately in the ratio:
utility function. P P 0 r (1 a)
= .
P N 1 + 0 a + r (1 a)
5.1. Partially covered market This is increasing with 0 and decreasing with N. Hence, the range
increases when the treatments become less useful either because
This section extends the previous analysis to the case where the option of no treatment becomes less detrimental or because
individuals may forgo any treatment. We are particularly interested the diversity of drugs decreases. The range also increases with the
58 D. Bardey et al. / Journal of Health Economics 47 (2016) 5063

The equilibrium number of rms is thus lower when the opportu-


nity benet of forgoing a treatment is high. Using this zero prot
free entry condition, the actuarial premium must just cover the
total cost of entry so that one has  = (1 + )Nk. It follows that the
socially optimal level of r is obtained by maximizing welfare (15)
subject to wh = w (1 + )kN and condition (18).

Corollary 1. Suppose that the copayment rate a = 0, then the optimal


diversity N is decreasing with the opportunity benet of forgoing a
treatment 0 .

Proof. See Appendix H.

As the opportunity cost of forgoing a treatment decreases, entry


becomes less desirable. Note that all the comparative statics in
Proposition 4 still hold.

Fig. 2. Price equilibria with elastic demand. 5.2. Separable utility function

strength of reference pricing r. Fig. 2 illustrates the price equilibria Until now, we have used a particular specication of the prefer-
in the (r, P) space. Every price between the upper and lower curves ences against side effects. It may be asked whether our main result,
are possible price equilibria in the market with an elastic demand. namely that less than maximal reference pricing is optimal, still
We now turn to the welfare analysis. In this model, we have holds under an alternative model. To show this, we now use a
to consider three situations for individuals. An individual may be model in which the side effect of a drug enters separately in the
healthy, unwell with treatment or unwell without treatment. This utility function. Formally, we now assume that the policyholder js
implies that we adjust our notion of index of diversity and that we expected utility when consuming drug i is:
introduce a WID that accounts for the probability of not using a     
U =  u wsi v xji + (1 )u(wh ),
treatment. Considering that policyholders preferences are repre-
sented by a CARA utility function and that Assumption 1 holds, one  
can rewrite the expected utility function as:
where v (x) = exp x and  represents the absolute risk aver-
1 1
 sion parameter with respect to adverse effects. Keeping the
EU =  u (ws ) + u (wh ) + (1 ) u (wh ) , (15) assumption that xji follows an i.i.d. exponential distribution, the
Q1 Q0
expected utility function now writes:24
where:
   
1 N 0 aP 1
= exp , (16) EU =  u (ws ) + (1 )u(wh ),
Q1 N + 0    (N)
Q
1 0
 NaP 
= exp . (17) at a symmetric equilibrium where
Q0 N + 0  
Q1 is the WID accounting both for the probability of consuming  (N) = 1 
Q .
a drug and the expected side effect conditional on the fact that N
individuals consume a drug. This WID depends on prices as this Following the same reasoning as in Section 3.1, we can state the
affects the decision to consume. In contrast, Q0 measures the value following proposition:
of the new option of not consuming drugs when the severity of ill-
ness is low. Contrary to Q1 , it is obviously decreasing in 0 since it Proposition 8. For a given copayment rate a, there exists a reference
decreases the disutility of not consuming any drug. As in the pre- pricing strength r inducing a free entry equilibrium with N drugs if
ceding section, both Q1 and Q0 are decreasing (resp. increasing) in
kN (N 1) N
 kN

the risk aversion parameter (the treatment diversity).   exp  ( (1 + ) (1 a) + a)

Notice that both Q1 and Q0 are functions of aP. This reects the  N + 
effect of the out-of-pocket payment on the probability of using
a treatment and on the conditional expected utility. Indeed as p
 exp 
(w) exp (w)
, ;
increases, it is less likely that the patient uses a treatment. This in  a
turn also implies that a patient using a treatment supports lower
side effects on average. the price is then P = kN/.
To conduct our policy analysis we need to make an assumption
Proof. See Appendix I.
on the relevant price that emerges given the policy mix (a, r). Given
that we are interested by, among other things, the conditions under The long run equilibrium is more complex than that described
which the optimal policy involves r < 1 under some insurance a < 1, in Eq. (8). This is due to the fact that an increase in the price paid by
for conciseness, we develop the analysis only for a = 0 and P = P, the patients now has an income effect: increasing the price decreases
largest possible equilibrium price. Then, we have P = / (N 1) r the disposable income when unwell which increases the marginal
and the average demand for a drug is D = / (N + 0 ). In the long utility of income. The semi-elasticity of residual demand is there-
run, the free-entry condition k = PD/N yields the diversity as an fore higher with this specication of preferences.
implicit function r as follows:


(N + 0 ) (N 1) r = . (18)
k 24
We assume that  < N, which is the counterpart of Assumption 1.
D. Bardey et al. / Journal of Health Economics 47 (2016) 5063 59

The problem of the insurer thus now amounts to solve: introducing sequential entry and delays in order to take into
  account the impact of health insurance and drug price regulation
1 on drugs introduction delay. Indeed, these delays of introduction
maxEU = (1 )u (wh ) +  u (ws ) ,
N,a  (N)
Q may have a huge impact in terms of welfare. All these aspects are
kN left for future research.
s.t. wh = w (1 + )(1 a)kN, ws = wh a ,

  Appendix A. Proof of Proposition 1
kN (N 1) kN exp (w)
exp  ( (1 + ) (1 a) + a) (r 1),
  
  The equilibrium condition is
kN (N 1) kN exp (w)
exp  ( (1 + ) (1 a) + a) (r 0).
    D
+ (a + (1 a)r) P = 0. (19)
N pi
Following the same reasoning as in Section 4, we can state the
following: Given prices pi and p, the probability that drug i is chosen by a
patient j is determined by
Proposition 9. Assuming an interior solution for a* , the optimal
policy mix veries r* < 1 if  1.
  
Pr pi + xji = min p +  xjl | xji
l
Proof. See Appendix J.  
l pi p
One can see that the main result still holds with this kind of = Pr min xj + xji | xji
l=
/ j 
preference. Everything works as if the risk aversion parameter with
 p p N1
respect to side effects risks  corresponds to the product of the i
= 1F + xji .
risk aversion parameter and the side effects marginal loss  in 
the previous model. A noticeable difference here is that  1 is a
sufcient (but not necessary) condition to have less than maximal Integrating over all possible realizations of xji , the expected demand
reference pricing. of drug i is thus:

 p p N1
i
6. Conclusive remarks Di =  f (x) 1 F +x dx
0


Our analysis sheds light on various mechanisms and underlines  N1


 
= exp (pi p) .
several results related to drug price regulation, such as reference N 
pricing. First, in the short run, it is always optimal for the health
It yields
insurer to implement the strongest policy in order to lower health
care expenditure. In the long run, the regulator takes into account D 
  N1
the drug entry process and the overall quality offered to patients. = .
pi N 
Diversity of treatment is an indirect form of insurance that may
substitute for or complement monetary insurance. In terms of pol- Substituting this expression in (19) thus yields:
icy making, our model points out the importance of having good 1 
estimations of the impact of prices on the supply and of the social P= .
a + (1 a)r N 1
value of drugs diversity (which depends positively on the level of
policyholders risk aversion) before deciding on a reference pricing
rule. Appendix B. Proof of Proposition 2
Our model focused on a specic price regulation mechanism,
namely reference pricing. Our main message would not change if For a xed price P = (N) the slope EU/a increases in
we introduce other price mechanism controls such as price caps or implying that the optimal copayment is non-decreasing in . The
exogenous reference pricing as it is done in Europe. The important rst order condition at a = 0 can be rewritten (using wh = ws =
aspect of the analysis is to show that a lenient price control can w (1 + )P)
be optimal in a context where drugs diversity is valued by risk
averse patients. One important and neglected aspect of our analysis  u (w )
(1 ) u (w ) +  u (w ) ,
is that we do not model vertical differentiation so that one can Q (N) Q (N) (1 + )
not see how the policy may affect breakthrough innovations. As or
suggested by Brekke et al. (2007), this would allow us to compare
 Q (N)
more explicitly the two modalities of reference pricing, i.e. generic (1 ) . (20)
(1 ) Q (N) + 
versus therapeutic. This would also be useful to take into account
the situation where brand-name drugs and generic drugs compete Similarly, we have a = 1 if
in the same reference pricing class.25 We however expect that long-
 u (w P)
term entry considerations would still mitigate the benets of tough (1 ) u (w) +  u (w P) ,
price regulation as long as reference pricing hurts the prots of all
Q (N) Q (N) (1 + )
active producers in the market. or
Finally, we have determined the optimal public regulation in  exp ( (N)) Q (N)
short and long run equilibrium analysis. Nevertheless, as in Bardey (1 ) . (21)
(1 ) Q (N) +  exp ( (N))
et al. (2010), it would be useful to complete this analysis by
The RHS of (20) and (21) decrease in N and take value 0 when Q = 1.
Hence, (20) holds for all Q if 0 . When (1 ) / > > 0 on
25
This may occur when dates of patent rights expiration differ within a given
the contrary, the condition (21) is satised if N N where N solves
therapeutic class, or when the sole brand-name drug stays in the market after patent (21) with equality. Finally, when (1 ) /, the condition holds
expiration. for all N.
60 D. Bardey et al. / Journal of Health Economics 47 (2016) 5063

Appendix C. Proof of Proposition 3


  
Hence, we have 2/  N  + k (2a + (1 a) (1 + )) < 0
so that the rst term in brackets in the RHS of (23) is nega-
We rst prove that the objective is quasi-concave in Lemma 1 tive hence /N|=0 < 0 at any point where  = 0. This implies
and then prove Proposition 3. that 2 EU/N2 < 0 when EU/N = 0 so that the objective is quasi-
Lemma 1. The objective is quasi-concave in N. concave.

Proof. The slope By quasi-concavity, r < 1 if and only if the RHS of (22) is positive

EU
 i.e. if:
=  u(ws ) k (1 + ) (1 a)   Q (N) 
N N Q k
  
  Q 2 N Q (N)
(1 ) u (wh ) + u (ws ) aku (ws )  kN

Q Q
(1 ) exp a
can be written as u (ws )  where  
(1 a) k (1 + ) Q (N) > 0
 
  +
= k (a + (1 a) (1 + )) k(1 + ) Q (N)
 N Q Q (24)
kN
 
(1 a) (1 ) exp a . (22) where the LHS of (24) is concave in a. This implies that r < 1 on an

interval (a, a).
We have One has
 


= k (a + (1 a) (1 + )) +  Q  
N  N 2 Q N Q =   ,
Q 2 N N N  Q
k
 kN
  
a (1 a) k (1 ) exp a (1 + ) . so that
 
 
Hence:   Q (N)   k 
k = N +    .
 

  Q (N)2 N Q (N) kN N  Q (N)
|=0 = k (a + (1 a) (1 + )) +
N  N 2 Q N Q But the free-entry condition yields
  
k    
a k (a + (1 a) (1 + )) . kN = 1 = N.
  N Q Q (N 1) kN
Using Therefore, we obtain

     k
  Q (N)  
Q
=  2 =  ,  k = 1 +  .
N N  N N  Q  Q (N) 2 N Q (N) N  Q (N)

 2 2 
= Substituting this last expression in (24) it yields r < 1 if
 3 =  2 Q ,
N 2 Q N  N N     1 a  EU
 1 
k > |N , (25)
we obtain N  Q (N) Nu (ws ) a
  2  
|=0 =
  2 Q + k (a + (1 a) (1 + ))   + where EU |N is evaluated at r = 1.
N N N  N N  Q a
           
1a EU 1a (1 ) u (wh )  
k  |N = +  (1 + )  N
a     k (a + (1 a) (1 + )) Q , Nu (ws ) a N u (ws ) Q Q
  N N  Q           
1a  
= (1 ) exp a N +  (1 + )  N ,
so N Q Q
2

Q     
which is decreasing with a when positive.
| =  N  Suppose that  > 1. Then as 1a EU | decreases when
 N =0 Nu (ws ) a N
+k (2a + (1 a) (1 + )) positive, condition (25) holds on an interval (a, 1) . This interval is
 k non-empty as the RHS vanishes at a = 1. At a = 0, the condition is
  a k (a + (1 a) (1 + )) . (23) veried if EU/a|N,a=0 < 0 (i.e. full insurance is optimal), in which
N N  
case a = 0.
Notice that  = 0 implies that 
For  = 1, the same conclusion  holds because our assumption
 that N < N ensures that 1a EU | <0 for a close to 1.
  Nu (ws ) a N
k (a + (1 a) (1 + )) =
 N Q Q  If <1,EUthe condition is violated at a
1. The function
  1a
|N is convex. Thus, the condition holds on some inter-
Nu (ws )
  a
  k (a + (1 a) (1 + )) >0 val (a, a) which may be empty.
N N  Q

implying (because N > ): Appendix D. Proof of Proposition 4



k (a + (1 a) (1 + )) <  . At any point where Na is interior, the effect of a parameter on
 N 
Na is given by the sign of the derivative of  with respect to this
D. Bardey et al. / Journal of Health Economics 47 (2016) 5063 61

parameter. Differentiating the LHS of (22) with respect to a and because a > 0 and N = 0. But this contradicts the assumption
respectively yields: made that a* decreases in . Thus, a* is non-decreasing with and
a* = 0 for . Notice also that at an interior solution

 1
 EU
= |N
dN
a Nu (ws ) a = 0.
kN kN
  d
+ (1 a) k (1 ) exp a (1 + ), Hence, N* is constant in when it is larger than N on > . Since
 
    N* decreases in for from Proposition 4, we conclude that
 kN  notice that N (N) =
= (1 a) k (1 ) exp a + < 0.  everywhere.
it is non-increasing   Moreover,

 Q kN (N 1) kN N  = kN  1 is positive if and only if
 > 1. Hence r* < 1 for all a* if  > 1. Finally, if  < 1, r* = 1 if a* > 0.
so that /a > 0 if EU/a|N > 0 and / < 0. Then, using


   Appendix F. Proof of Proposition 6
=  
N Q N N  Q
First, from Eq. (27), it is immediate that N* increases in ,  and
we have , and N* decreases in k . Applying the Cramers rule yields

Q  a aN
 =   (a + (1 a) (1 + )) k
| |
 N N  da N NN
  N  =
kN d det
(1 a) k (1 ) exp a (1 + )
 N where


which is increasing in  and . Thus / > 0 and / > 0 at kN 
a = (1 + ) (1 ) exp a (1 + a )
=0.  N


kN 
Appendix E. Proof of Proposition 5 aN = (1 + ) (1 ) exp a (1 a )
 N 2

Suppose a* decreases in for some value. Then it must be the N = kN


case that a* > 0 . Moreover N > N because a* is non-decreasing in  
NN = k 2N 
for a xed N from Proposition 2. At such a point, the rst-order
conditions are: so that
 

u (ws ) da
2 (1 + ) (1 ) k exp a kN


Q = N  + a N < 0.
EU (.) = 0, d N det
1+

 
Moreover,
 u(ws ) k u (ws ) = 0,
N Q Q a aN
| |

which can be rewritten (up to a proportional factor) as: da N NN
=
    d det
N  kN

a = (1 + ) (1 ) exp a where
N  

 kN
+  (1 + ) 1 = 0, (26) a = (1 + ) (1 ) exp a
N 

N =  + kN 

 
N =  kN N  = 0. (27) so that
 
  da
 (1 + ) (1 ) exp a kN

We have at a = N = 0 (using N  k/ = /N for aN ): =
d N det
   
  k kN
  
 
aa = (1 + ) (1 ) N   exp a < 0, k 2N  +  + kN  (1 a ) .
  N


 
kN  Using  = kN N  , we obtain
aN = (1 + ) (1 ) exp a (1 a ) > 0,
 N 2  
da
 (1 + ) (1 ) exp a kN

 
Na = 0, NN = 2kN + k < 0. = k N (1 + a )  < 0.
d N det
First, notice that the determinant of the modied Hessian is Finally, we have
det = aa NN > 0, hence the matrix is semi-denite negative. Dif-
ferentiation of a* with respect to gives ak aN
| |

da Nk NN
a aN = < 0,
| | dk det
a N NN a NN as ak , NN and Nk are negative while aN is positive. Similarly,
= = > 0,
det det a < 0 and N = 0 imply da* /d < 0.
62 D. Bardey et al. / Journal of Health Economics 47 (2016) 5063

Appendix G. Proof of Proposition 7 Given prices pi and p, the probability that drug i is chosen by a
patient j is determined by
Since     
i max
Pr u wsi exp x u (ws ) exp  xjl | xji
j
l
Pi arg max Pi Di (Pi , P)
  
 
every price Pi that satisfy = Pr min exp  xjl i | xi
u (ws ) u wsi + exp xj j
l=
/ j

Di (Pi , P)     N1
Di (Pi , P) + Pi |Pi P 0for Pi = P i
ln u (ws ) u wsi + exp x
Pi j
= 1 F .
Di (Pi , P) 
Di (Pi , P) + Pi |Pi P 0for Pi = P
Pi
Integrating over all possible realizations of xji , the expected demand
are candidates for symmetric equilibria. Since of drug i is thus:

      N1
Di (P, P)
ln u (ws ) u wsi
+ exp x
Di + Pi |Pi P = 0 Di =  f (x) 1F dx
Pi 
0

leads to
N1

   
=  expx u (ws ) u wsi + exp x
 dx

Pi = 0
(N 1) + 0 a
Differentiating Di with respect to pi yields:
and Di  
=  (N 1) u wsi
Di (Pi , P) pi
Di + Pi |Pi P = 0
  N 1 1
Pi
 i
exp x
u (ws ) u ws + exp x  dx
leads to 0

 
 
x N+
 =  (N 1) u wsi exp dx
Pi =
(N 1) + 0 0
  1
  =  (N 1) u wsi  
and that Pi < Pi , any price belonging to the interval Pi , Pi = N + 
 
P, P are possible candidates for the symmetric price competition
game. so that the symmetric equilibrium price is given by:
 
N + 
Appendix H. Proof of Corollary 1 1
P exp [ ( (1 + ) (1 a) + a) P] exp (w) =
N (N 1)
The optimal diversity N solves
The long term equilibrium can thus be described by:
 N + 0
  
maxEU = u (wh ) + (1 ) u (wh ) N + 
N N + 0  1 exp (w)
P exp [ ( (1 + ) (1 a) + a) P] =
 (N 1) N 
s.t. wh = w (1 + )kN and (N + 0 ) (N 1)
k P = kN
The gain of raising diversity is now captured by: which yields:
   
kN kN (N 1) N exp (w)
EU  exp  ( (1 + ) (1 a) + a)   =
=   2 u (wh )   
N N+
N + 0 
 N + 0
 (28)
(1 + ) k (1 ) +  u (wh ) .
N + 0  It is easy to show that N is decreasing in [a, 1] which proves
Proposition 8.
Thus at any interior equilibrium one has
   Appendix J. Proof of Proposition 9
 (1 + ) k N + 0  N + 0 (1 )  = 0.

The LHS is decreasing with N and 0 , which shows that N is unique Assuming an interior solution for a one has:
and N/ 0 < 0 when N is interior. EU
= kN (1 + ) EU kNu (ws ) = 0. (29)
a
Appendix I. Proof of Proposition 8 Differentiating the objective function with respect to N yields:

The rst order condition for prot maximization of rm i is still EU Q


=  k (1 + ) (1 a) EU aku (ws ) . (30)
given by (19). N N
D. Bardey et al. / Journal of Health Economics 47 (2016) 5063 63

Substituting (29) in (30), the gain from rising diversity above N is Brekke, K., Konigbauer, I., Straume, O.R., 2007. Reference pricing of pharmaceuticals.
given by: Journal of Health Economics 26, 613642.
Brekke, K., Grasdal, A.L., Holms, T.H., 2009. Regulation and pricing of pharmaceut-
EU Q icals: reference pricing or price cap regulation? European Economic Review 53,
|N =  ku (ws ) (31) 170185.
N N Brekke, K., Holms, T.H., Straume, O.R., 2011. Reference pricing, competition, and
pharmaceutical expenditures: theory and evidence from a natural experiment.
where N is implicitly given by (28) where = 1:
Journal of Public Economics 95, 624638.
N 2 (N 1)
 kN
 Danzon, P.M., 2001. Reference Pricing: Theory and Evidence. In: Lopez-Casasnovas,
  k exp  ( (1 + ) (1 a) + a) = exp (w) (32) G., Jonsson, B. (Eds.), Reference Pricing and Pharmaceutical Policy. Springer, New
 York.
 N + 
Danzon, P.M., 2012. Regulation of Price and Reimbursement for Pharmaceuticals,
The Oxford Handbook of The Economics of the Biopharmaceutical Industry.
Substituting (32) in (31) thus yields: Oxford University press, pp. 266301.
  Danzon, P., Ketchman, J., 2004. Reference pricing of pharmaceuticals for medicare:
evidence from Germany, the Netherlands and New Zealand. Frontiers in Health
N + 
EU  Policy Research 7 (2), bepress.
|N =   2 Di Masi, J.A., Paquette, C., 2004. The economics of follow-on drug research and
N N 2 (N 1) development. Pharmacoeconomics 22 (Supl 2), 114.
N  Finkelstein, A., Luttmer, E., Notowidigdo, M., 2013. What good is wealth without
   2   health? The effect of health on the marginal utility of consumption. Journal of
the European Economic Association 11, 221258.
N 3  1 +  N  Frothingham, R., 2004. Me-too products friend or foe? New England Journal of
=  2 Medicine 350 (20), 21002101.
Geoffard, P.-Y., 2006. Incentive and selection effects in health insurance. In: Jones,
N  N 2 (N 1)
A. (Ed.), Elgar Companion to Health Economics. Elgar.
Lakdawalla, D., Sood, N., 2009. Innovation and the welfare effects of public drug
which is positive if  1. insurance. Journal of Public Economics 93 (3-4), 541548.
Lakdawalla, D., Sood, N., 2013. Health insurance as a two-part pricing contract.
Journal of Public Economics 102, 112.
References Lopez-Casanovas, G., Puig-Junoy, J., 2000. Review of the literature on reference pri-
cing. Health Policy 54, 87123.
Anderson, S.P., de Palma, A., Nesterov, Y., 1995. Oligopolistic competition and the Miraldo, M., 2007. Reference Pricing Versus Co-Payment in the Pharmaceutical
optimal provision of products. Econometrica 63 (6), 12811301. Industry: Firms Pricing Strategies. Center for Health Economics, University of
Bardey, D., Bommier, A., Jullien, B., 2010. Retail price regulation and innovation: York, Research Paper 27.
reference pricing in the pharmaceutical industry. Journal of Health Economics Ruggeri, K., Nolte, E., 2013. Pharmaceutical Pricing: The Use of Reference Pricing.
29 (2), 303316. Rand Corporation Research Report Series.
Bardey, D., Jullien, B., Lozachmeur, J.-M., 2013. Health Insurance and Diversity of Wertheimer, A., Levy, R., OConnor, T., 2001. Too many drugs? The clinical and eco-
Treatment: A Policy Mix Perspective. In: CESifo Working Paper Series No. 4111. nomic value of incremental innovations. In: Irina Farquhar, Kent Summers, Alan
Brekke, K., Canta, C., Straume, O.R., 2015a. Reference pricing with endogenous Sorkin (Eds.), Investing in Health: The Social and Economic Benets of Health
generic entry. In: NHH Working Paper. Care Innovation (Research in Human Capital and Development, Volume 14).
Brekke, K., Canta, C., Straume, O.R., 2015b. Does Reference Pricing Drive Out Generic Emerald Group Publishing Limited, pp. 77118.
Competition in Pharmaceutical Markets? Evidence from a Policy Reform. In:
NIPE Working Paper.

You might also like