Hostname: page-component-cd9895bd7-gbm5v Total loading time: 0 Render date: 2024-12-26T10:28:28.316Z Has data issue: false hasContentIssue false

Insurance design and arson-type risks

Published online by Cambridge University Press:  26 November 2024

Jean-Gabriel Lauzier*
Affiliation:
Department of Economics, Memorial University of Newfoundland, St. John’s, Newfoundland and Labrador, Canada
*
Rights & Permissions [Opens in a new window]

Abstract

We design the insurance contract when the insurer faces arson-type risks. We show that the optimal contract must be manipulation-proof. Therefore, it is continuous, has a bounded slope, and satisfies the no-sabotage condition when arson-type actions are free. Any contract that mixes a deductible, coinsurance, and an upper limit is manipulation-proof. A key feature of our models is that we provide a simple, general, and entirely elementary proof of manipulation-proofness that is easily adapted to different settings. We also show that the ability to perform arson-type actions reduces the insured’s welfare as less coverage is offered in equilibrium.

Type
Original Research Paper
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2024. Published by Cambridge University Press on behalf of Institute and Faculty of Actuaries

1. Introduction

Suppose Bob was just involved in a bicycle accident. After the fact, an officer of the law provided Bob with a certificate indicating that the automobile driver was responsible for the accident. However, the certificate does not specify how bad the damage inflicted on Bob’s steel steed was. Which types of insurance contracts will incentivize Bob to take a sledgehammer to his bicycle before taking a picture and filing his insurance claim?

An arson-type action is the action of inflating an insurance claim by physically destroying an object without being caught. In this article, we show that the optimal contract is manipulation-proof in the sense that it never incentivizes the insured to perform arson-type actions. Manipulation-proofness requires the optimal contract to be continuous and to have a bounded slope. Our model thus implies the no-sabotage condition of Carlier & Dana (Reference Dana2003) when increasing the damage is free (Bob already has a sledgehammer).

We provide a simple model that yields a simple, general, and entirely elementary proof of the optimality of manipulation-proof contracts. A key advantage of our proof is that it does not rely either on the Revelation Principle (Myerson, Reference Myerson1981) or on the properties of the agents’ decision criteria. It is, therefore, easy to adapt to different settings. Our model is thus quite general and nests previous results, as we explain below in Section 1.1.

The argument proceeds in two steps. We first show that the value function of the manipulation stage of the game is a Lipschitz continuous function, with a Lipschitz constant proportional to the cost of inflating the loss. We then show that for every contract inducing arson-type actions, an alternative contract provides realization-by-realization of the same coverage and does not induce manipulations. This new contract is precisely equal to the aforementioned value function. Contracts with arson-type risks are priced at a higher premium, and thus, the latter contract is cheaper than the former and dominates it. Insurers only offer manipulation-proof contracts in equilibrium. As such, the ability to perform arson-type actions hurts the insured because it strictly reduces the amount of coverage offered in equilibrium. We highlight that the types of contracts frequently observed in real-world insurance markets are robust to arson-type risks. These include all contracts that mix deductible, coinsurance, and an upper limit.

Intuitively, Bob will never want to pay for a protection that lets him destroy his bike and profit. Bob’s protection is thus somewhat limited, to the extent that he would benefit from committing not to buy a sledgehammer. Of course, Bob is not credible and is only offered contracts with limited protection. Since all contracts mixing deductibles, coinsurance, and upper limits are manipulation-proof, we can assume as a rule of thumb (heuristic) that that’s what Bob will purchase.

Next, we discuss the literature. Section 2 serves as the core of the paper. Section 2.1 introduces a general insurance design model with arson-type risks and shows that the optimal contract must be manipulation-proof. This yields a general representation of the optimal contract, which we then use in Section 2.2 to provide closed-form solutions of the optimal contract when the insurer faces a variety of different administrative costs to process the claim. Section 3 concludes by discussing the model’s robustness and its interest as a heuristic for risk professionals.

1.1 Related literature

It is well-known that the possibility of inflating an insurance claim by physically destroying an object imposes structure on the type of contract that an insurer can offer. Huberman et al. (Reference Huberman, Mayers and Smith1983) contain the earliest mention of arson-type actions we are aware of. The authors analyze the optimal insurance contract when respecting the contract involves administrative costs but where there are economies of scale. Their model’s optimal first-best contract is a completely disappearing deductible (Fig. 1a). They show that if the insured can cause extra damage, then the second-best contract is a straight deductible ( Fig. 1b). Similarly, Picard (Reference Picard2000) introduces arson-type actions to restrict the set of feasible contracts in a setting where the insured can defraud the contract and manipulate the audit costs. The model’s first-best contract is discontinuous, and the author shows that this discontinuity disappears when there are arson-type risks.

Figure 1 Disappearing v.s. straight deductible.

Following the articles mentioned above, it has become routine in the literature to assume that the retention schedule is monotonic. For instance, Cai & Chi (Reference Cai and Chi2020) ’s survey contains an exhaustive overview of the use of this (co)monotonicity assumption in the reinsurance literature. This assumption is now referred to as the no-sabotage condition as of Carlier & Dana (Reference Dana2003). The name refers to the observation that non-monotonic retention schedules incite the insured to inflate its loss.

Our model provides a simple, formal counterpart to such an argument. We obtain the contract’s continuity, its bounded slope, and the no-sabotage condition as implications of the same result. Specifically, the optimal contract is always $1$ Lipschitz when the arson-type actions are free, the case of our $\beta$ parameter being equal to zero. The proof is simple and “robust” in the sense that it does not rely on many properties of the underlying optimization problem and is thus easily adaptable to different settings. In this sense, although our argument is different, our model nests many previous arguments used to derive no-sabotage, including the classic arguments of both Huberman et al. (Reference Huberman, Mayers and Smith1983) and Picard (Reference Picard2000).Footnote 1 At a technical level, we borrow from the Costly State Falsification literature (Crocker & Morgan, Reference Crocker and Morgan1998; Lacker & Weinberg, Reference Lacker and Weinberg1989; Maggi & Rodriguez-Clare, Reference Maggi and Rodriguez-Clare1995) to model arson-type actions as ex-post hidden actions for which the “cost” of manipulating the observed losses is “proportionally high.” We discuss the topic further in Section 3.

An essential part of our motivation is to provide a simple proof that sheds light on the structure of the no-sabotage condition. The argument we provide is easily seen to hold for different contracting games and does not invoke the revelation principle (Myerson, Reference Myerson1981). We surmise that our model can be interpreted as a micro-foundation for the heuristic of using “simple” comonotone contracts. In other words, the article complements the well-established risk-sharing literature with comonotone contracts (Boonen et al., Reference Boonen, Liu and Wang2021; Carlier et al., Reference Carlier, Dana and Galichon2012; Dana & Meilijson, Reference Dana and Meilijson2003; Landsberger & Meilijson, Reference Landsberger and Meilijson1994; Ludkovski & Rüschendorf, Reference Ludkovski and Rüschendorf2008) by showing how comonotonicity also naturally obtains as the equilibrium outcome of contracting games with arson-type actions.

Many types of contracts used in practice satisfy the no-sabotage condition and are, therefore, manipulation-proof with regard to arson-type actions. This includes full insurance contracts, straight deductibles, and pure coinsurance contracts. Thus, if a contract is the first-best solution to a problem of insurance design without arson-type actions, it is also the second-best solution to the same problem with arson-type actions. In other words, the possibility of arson-type actions affects the design of insurance contracts if and only if the first-best contract is not itself manipulation-proof. Such departures routinely happen when the insurer faces administrative costs to respect the contract. We, therefore, cast our model in the context of both fixed (Spaeter & Roger, Reference Spaeter and Roger1997), which contains a general analysis of the design of insurance contracts with administrative costs. This approach, combined with results on comonotone contracts derived in Carlier & Dana (Reference Carlier and Dana2005), allows us to streamline proofs and the exposition significantly.

2. Model

In Section 2.1, we introduce our baseline model and prove our main result, the manipulation-proofness of the optimal contract. This implies that the optimal contract is continuous and has a bounded slope, and we then characterize the optimal contract as a generalized deductible. We turn to applications in Section 2.2, where we obtain closed-form solutions of the optimal contract in a variety of settings. We first establish that arson-type risks are irrelevant in the benchmark Arrow-Borch-Raviv model, where there are no administrative costs. We then study the case of fixed administrative costs to process claims as in Gollier (Reference Gollier1987). Through simple arguments, we show that the ability to use arson-type actions strictly hurts the insured. This is because, in equilibrium, the optimal insurance contract provides less coverage than it would otherwise. We close this section with an analysis of the second-best contract when the first-best is a completely disappearing deductible as in Huberman et al. (Reference Huberman, Mayers and Smith1983).

2.1 Baseline model and manipulation-proofness

Let $\mathbb{P}$ be a probability measure and let the risk $X$ be a random variable that has some mass at $0$ and is continuous on $(0, M]$ , where it has full support. The probability mass is the probability that no accident occurs. Let the function $c\,:\,[0, M] \rightarrow \mathbb{R}_+$ be the administrative cost of respecting the insurance contract and let $Y\,:\,[0, M]\rightarrow [0, M]$ be the indemnity schedule.Footnote 2 The amount $H\geq 0$ denotes the price (premium) of the contract while $W_0\gt 0$ is the initial wealth of the insured and $\rho \geq 0$ is the loading factor. As usual, the function $u\,:\,\mathbb{R}_+ \rightarrow \mathbb{R}_+$ is a twice-differentiable and strictly concave Bernoulli utility function satisfying Inada conditions.

The game proceeds as follows:

  1. Stage 1 the insured buys the insurance contract $Y$ at price $H$ ;

  2. Stage 2 the loss $x\in [0,M]$ occurs (Nature moves);

  3. Stage 3 the insured observes the loss and decides to take hidden action $z\in [0, M - x]$ to augment the damages andFootnote 3

  4. Stage 4 the contract is implemented without renegotiation.

The insured’s terminal wealth is therefore

\begin{equation*}W\,:\!=\,W_0 - H - x - z -g(z)+Y(x+z). \end{equation*}

The solution concept is a weak Perfect Bayesian equilibrium (Mas-Colell et al., Reference Mas-Colell, Whinston and Green1995) where we assume that the insured takes the insurer’s favored action whenever indifferent.

By backward induction, the optimal contract solves the program:

(Problem I) \begin{align} \sup _{H \geq 0, Y \in B_+(\mathcal{B}([0,M]))} & \int u(W_0 - H - x - z^*(x) -g(z^*(x))+Y(x+z^*(x)))\,\textrm{d} \mathbb{P}\\[-4pt]\nonumber \end{align}
(LL) \begin{align} \textrm{s.t.}\, & \,0 \leq Y \\[-4pt]\nonumber \end{align}
(B) \begin{align} &\, \forall x,\, Y(x + z^*(x)) \leq x +z^*(x) \\[-4pt]\nonumber \end{align}
(PC) \begin{align} & (1+\rho )\int Y(x+z^*(x))+c(Y(x+z^*(x))) \,\textrm{d} \mathbb{P} \leq H \\[-4pt] \nonumber\end{align}
(IC) \begin{align} & \forall x\, z^*(x) \in \arg \max _z \left\{ Y(x+z)- z -g(z) \right \} \\[10pt] \nonumber\end{align}

where (LL) is the insured’s limited liability constraint, (B) is the boundedness constraint stating that the insurer will never pay more than the observed loss, (PC) is the insurer’s participation constraint, (IC) is the insured’s incentive compatibility constraint and the function

\begin{align*} g(z)= \begin{cases} +\infty &\text{ if } z\lt 0\\[3pt] \beta z &\text{ if } z\geq 0 \end{cases} \end{align*}

represents an extra cost of inflicting damage (Bob buying a sledgehammer).

Let us start with a handy assumption which is common in the literature:

Assumption 1. We assume throughout that feasible contracts are non-decreasing and upper semi-continuous functions: for every $x_0\in [0, M]$ it is

\begin{equation*}\limsup _{x\rightarrow x_0} Y(x)=Y(x_0).\end{equation*}

Of course, this implies that the optimal contract $Y$ is non-decreasing and upper semi-continuous. Assumption1 is standard and is not discussed further. The next statement is standard and will not be proved:

Lemma 1. The insurer’s participation constraint ( PC ) must be binding.

When the loading factor $\rho$ is zero, then Lemma1 states that the contract must be sold at an actuarially fair price.

Let us now inspect the incentive compatibility constraint (IC). This constraint is an optimization program which is ill-behaved. This is because we do not know at this level of generality if the optimal contract $Y$ is continuous. Moreover, as discussed in Section 1.1, we cannot assume away the possibility that $Y$ is discontinuous. We seem to be in trouble now because we cannot differentiate the objective function in (IC) and thus cannot use standard first-order conditions to characterize the (set of) optimal manipulations. Fortunately, the optimization problem of constraint (IC) is more straightforward to characterize than it looks at first sight.

Let $V_Y\,:\,\mathbb{R} \rightarrow \mathbb{R}$ be the value function of the manipulation stage of the game so that

\begin{align*} V_Y(x)= Y\!\left(x+z^*(x)\right)- z^*(x) -g\!\left(z^*(x)\right) \end{align*}

for

\begin{align*} z^*(x) \in \sigma _Y(x)=\arg \max _z \left \{ Y(x+z)- z -g(z) \right \}, \end{align*}

where $\sigma _Y\,:\,[0, M]\rightrightarrows [0, M]$ denotes the optimal choice correspondence of the manipulation stage of the game induced by contract $Y$ .Footnote 4

Lemma 2. If $Y$ is non-decreasing and upper semi-continuous, then $V_Y$ is Lipschitz continuous with Lipschitz constant smaller or equal to $1+\beta$ .

Proof. We first show that $V_Y$ is continuous. Suppose, by contradiction, that there exists a $x_0\in (0, M]$ such that $\lim _{x\uparrow x_0}V_Y(x)\lt V_Y(x_0)$ . We can find an $\varepsilon \gt 0$ small such that for $x=x_0-\varepsilon$ we have $V_Y(x_0)-V_Y(x)\gt (1+\beta )\varepsilon$ . Letting $z^*(x_0)\in \sigma (x_0)$ and setting $z= z^*(x_0)+\varepsilon$ we obtain

\begin{align*} V_Y(x)&\geq Y(x +z)-z-g(z)\\ &=Y(x_0-\varepsilon +z^*(x_0)+\varepsilon )-z^*(x_0)-\varepsilon -g(z^*(x_0)+\varepsilon )\\ &=Y(x_0+z^*(x_0))-z^*(x_0)-\varepsilon - \beta (z^*(x_0)+\varepsilon )\\ &=V_Y(x_0)-(1+\beta )\varepsilon \\ &\gt V_Y(x), \end{align*}

an absurd. Lipschitzianity follows from the same argument assuming the contradiction hypothesis that there are $x,x^{\prime}\in [0, M]$ such that $\vert V_Y(x^{\prime})- V_Y(x)\vert \gt (1+\beta )\vert x^{\prime}-x\vert$ .

Fig. 2 depicts the value function $V_Y$ of the manipulation stage of the game for a given discontinuous contract $Y(x)=(x - \delta ){\unicode{x1D7D9}}_{\{x \geq b\}}$ , where $b \in (0, M)$ , $\delta \in (0, b)$ and $\beta \gt 0$ .Footnote 5 In the figure, manipulations always happen in the segment $(a,b)$ , where $a=(\delta +\beta b)/(1+\beta )$ . Clearly, we have

\begin{align*} V_Y(x)= \begin{cases} b-\delta -(1+\beta )x &\text{ if } x\in (a, b)\\[3pt] Y(x) &\text{otherwise}, \end{cases} \end{align*}

and the maximum Lipschitz constant of $V_Y$ is pinned down by $\beta$ .

Figure 2 The value function $V_Y$ of the manipulation stage of the game for $Y(x)=(x - \delta ){\unicode{x1D7D9}}_{\{x \geq b\}}$ for $b\in (0,M)$ , $\delta \in (0,b)$ and $\beta \gt 0$ .

Figure 3 Straight deductibles and constant retention contracts.

Since $V_Y \geq Y$ and $V_Y(x)\gt x$ implies that there is an $x^{\prime}$ such that $ Y(x^{\prime})\gt x^{\prime}$ , it is easy to prove that $V_Y(x)\geq x$ for some $x\in [0,M]$ if and only if $Y(x^{\prime})\geq x^{\prime}$ for some $x^{\prime}\in [0,M]$ . Thus, $V_Y$ satisfies all boundedness constraints whenever $Y$ does. The next Lemma states this observation formally; its proof is omitted since it is immediate.

Lemma 3. If $0\leq Y \leq X$ then $0\leq V_Y \leq X$ .

We are now ready to prove our main result. We say that a contract $Y$ is manipulation-proof if for every $x\in [0,M]$ it is $0 \in \sigma _Y(x)$ .

Theorem 1. Any optimal contract $Y$ is manipulation-proof.

Proof. Suppose, by contraposition, that $Y$ is optimal but that there exists a set $\chi \subset [0,M]$ such that simultaneously $\mathbb{P}[X\in \chi ]\gt 0$ and for every $x\in \chi$ it is $0\notin \sigma _Y(x)$ . Let $V_Y(x)$ be the value function of the manipulation problem defined by $Y$ , and consider now the alternative contract $(\overline{H}, \overline{Y})$ where $\overline{Y}=V_Y$ . That is, the new indemnity schedule gives realization-by-realization of the same final reimbursement as the original one after manipulations. Since $Y$ is optimal by assumption, it holds that $0\leq Y \leq X$ and $\overline{Y}$ satisfy all the boundedness constraints by Lemma3.

We now verify that the new indemnity schedule $\overline{Y}$ is manipulation-proof. By Assumption1 and Lemma2, $\overline{Y}$ is non-decreasing and Lipschitz with constant $1+\beta$ . By Lipschitzianity, for any two $x,x^{\prime}\in [0, M]$ such that $x\lt x^{\prime}$ we have

\begin{equation*}\overline {Y}(x^{\prime})-\overline {Y}(x) \leq (1+\beta )(x^{\prime}-x),\end{equation*}

or equivalently

\begin{equation*}\overline {Y}(x) \geq \overline {Y}(x^{\prime}) -(1+\beta )(x^{\prime}-x).\end{equation*}

The last inequality implies that for every realization $x\in [0,M]$ we have

\begin{equation*}0 \in \sigma _{\overline {Y}}(x)=\arg \max _z\left \{\overline {Y}(x+z) -z-g(z) \right \}.\end{equation*}

That is, $\overline{Y}$ does not induce manipulations and gives realization-by-realization of the same coverage as $Y$ .

The new contract strictly dominates the original one because

\begin{align*} \int Y\left (x+ z^*(x)\right )+c\left (Y\left (x+z^*(x)\right )\right ) \,\textrm{d} \mathbb{P} \gt \int \overline{Y}(x) + c\left (\overline{Y}(x)\right ) \,\textrm{d} {\mathbb{P}} \end{align*}

and the price $\overline{H}$ of $\overline{Y}$ is strictly smaller than the price $H$ of $Y$ .

We can understand Theorem1 as stating that since the insurer fully prices arson-type risks, contracts that induce arson-type actions will never be offered in equilibrium. Indeed, all extra damage due to arson-type actions must be priced at a higher premium. Of course, the insured will prefer the cheapest contract as he receives realization-by-realization of the same final amount under both contracts. Or, from Bob’s perspective, he was offered two contracts offering the same protection. An expensive one that allows him to take a sledgehammer to his bike and a cheaper one that does not. Bob, being a rational person, chooses the cheaper option.

Corollary 1. Any optimal contract $Y$ must be Lipschitz continuous with Lipschitz constant smaller or equal to $1+\beta$ .

Remark 1. Notice that the proofs leading Corollary 1 are entirely elementary, do not invoke the Revelation Principle, and do not rely on specific properties of the decision criterion used by any of the agents. It is clear that the arguments we used can easily be adapted to other settings and that we obtain the no-sabotage when $\beta =0$ . It is in this sense that our model “nests” classical no-sabotage results like those of Huberman et al. (Reference Huberman, Mayers and Smith1983)or Picard (Reference Picard2000).

Characterization.

Corollary1 implies that the family of contracts

\begin{equation*} \{Y\in B_+(\mathcal {B}([0,M]))\,:\,Y(x)=V_Y(x)\}\end{equation*}

consists of functions that are a.e. differentiable.Footnote 6 We can rewrite (Problem I) as

(Problem S) \begin{align} \max _{H \geq 0, Y \in B_+(\mathcal{B}([0,M]))} & \int u (W_0 - H - x+Y(x))\,\textrm{d} {\mathbb{P}} \\[-4pt] \nonumber \end{align}
(S1) \begin{align} \textrm{s.t.}\, & \,0 \leq Y \leq X \\[-4pt] \nonumber \end{align}
(S2) \begin{align} & \textrm{slope}(Y)\leq 1+ \beta \\[-4pt] \nonumber \end{align}
(S3) \begin{align} & (1+\rho )\int Y(x)+c(Y(x)) \,\textrm{d} \mathbb{P} = H \\[10pt] \nonumber\end{align}

under the implicit assumption that $Y\in C^0[0,M]$ is a.e. differentiable.Footnote 7 Notice how this problem, as rewritten, is almost identical to the problem studied in Spaeter & Roger (Reference Spaeter and Roger1997), except for the extra constraint $\textrm{slope}(Y)\leq 1+\beta$ .

We close this section with an immediate general representation theorem and a few observations that are handy when deriving closed-form solutions. As a consequence of Theorem1, any optimal contract $Y$ can be written as

\begin{align*} Y(x)=\max \{0, \alpha (x)x-d\}, \end{align*}

where $d\geq 0$ is a deductible and $\alpha (x)$ is a non-negative, continuous, and a.e. differentiable function satisfying for every $x\in [0, M]$

\begin{align*} 0 \leq \frac{\partial \alpha (x)x}{\partial x}\leq 1+\beta . \end{align*}

We say that the contract $Y$ is a full insurance contract when $d=0$ and $\alpha (x)=1$ everywhere; is a straight deductible when $d\gt 0$ and $\alpha (x)=1$ everywhere; entails coinsurance when $\alpha (x)=\alpha \in (0,1)$ ; and has upper limit $\gamma \gt 0$ if for every $x$ , $Y(x)\leq \gamma$ , with strict equality for some $x$ . Any contract mixing deductibles, coinsurance, and upper limits are manipulation-proof and can be written as

\begin{align*} Y(x)=\min \left\{\gamma, \max \{0, \alpha x - d\} \right\}. \end{align*}

We obtain the no-sabotage condition when $\beta =0$ . Formally, if $\beta =0$ then for any contract $Y$ the following holds:

  1. 1. $\textrm{slope(Y)} \leq 1$ ;

  2. 2. the retention function $R(x)=x-Y(x)$ is weakly monotone increasing;

  3. 3. the random variables $X,Y$ and $R(X)= X-Y(X)$ are comonotonic.Footnote 8

2.2 Some closed-form solutions

We now provide closed-form solutions to our contracting problem.

Full insurance, straight deductibles, and the irrelevance of arson

Lemma 4. [Arrow’s Theorem] If $c=0$ the optimal contract is a straight deductible $Y(x)=\max \{0,x-d\}$ for which $d=0$ if and only if $\rho =0$ .

Lemma4 need not be proved as it is the classic Arrow-Borch-Raviv Theorem (see Dionne et al., Reference Dionne2000). The lemma tells us that (Problem S) becomes interesting only when $c\gt 0$ somewhere. This is because when $c=0$ , the first-best contract never incentivizes Bob to take a sledgehammer to his bike. That is, arson-type risks impact the insurance only when there is a meaningful reason not to provide either full insurance or a straight deductible.

Fixed costs and nuisance claims: arson-type risks reduce welfare

We now consider the case when the administrative costs to deliver the contract are fixed costs. The setting with fixed cost was first introduced in Gollier (Reference Gollier1987) and is well-understood (Carlier & Dana, Reference Dana2003, Reference Carlier and Dana2005; Picard, Reference Picard2000, Reference Picard2013). We consider this case because it is the easiest way to show that arson-type risks hurt the insured. That is, to show that Bob would prefer to be unable to destroy his bike.

Assumption 2. The cost function involves only a fixed cost per claim: $c$ satisfies $c(0)=0$ and $c(y)=c_0\gt 0$ for every claim requiring the insurer to pay $y\gt 0$ .

We claim that when the insured can freely augment the damage ( $\beta =0$ ), then the optimal contract is a straight deductible.

Proposition 1. Under Assumption 2 , the solution to (Problem S) is a straight deductible: there exists a $d\gt 0$ such that

\begin{equation*}Y(x)= \max \{0, x-d\}.\end{equation*}

Proof. By Theorem1, our optimal contract $Y^*$ takes the form of a generalized deductible $Y^*(x)=\max \{0, \alpha (x)x-d\}$ . By Proposition 6 of Carlier & Dana (Reference Carlier and Dana2005), the optimal contract $Y^*$ is $1$ -Lipschitz on $\{Y\gt 0\}$ .Footnote 9 This implies that $\alpha (x)=1$ for all $x\in \{Y^*\gt 0\}=(d, M]$ , as desired.

We can now rewrite problem (Problem S) as

(Problem FC and A) \begin{align} \max _{d\geq 0} & \int u\! \left(W_0 - H - x +\max \{0, x-d\}\right)\,\textrm{d}\mathbb{P} \\ \textrm{s.t.}\quad & c_0 \mathbb{P}[X\geq d] + \mathbb{E}[X-d\vert X\geq d] = (1+\rho )^{-1}H \nonumber . \end{align}

We want to show that the possibility of arson-type risks hurts the insured, i.e., that Bob would like to commit to leaving his bike intact because he would be offered a better contract. Formally, we will show that absent the incentive compatibility constraint, a contract with an upward jump would improve upon a straight deductible. That is, we want to show that the straight deductible contract in Figure 3(a) is dominated by the discontinuous contract in Figure 3(b).

Once again, Proposition 6 of Carlier & Dana (Reference Carlier and Dana2005) implies that any optimal contract is $1$ -Lipschitz on $\{Y\gt 0\}$ . We can thus consider a contract of the form

\begin{align*} Y_{\delta, b}(x)=(x-\delta ){\unicode{x1D7D9}}_{\{x\geq b\}}=\begin{cases} 0 &\text{ if } x\lt b \\[3pt] x-\delta &\text{ if } x\geq b\end{cases} \end{align*}

for parameter $b\geq 0$ being a threshold, $\delta \geq 0$ being a loss retention parameter and $b \geq \delta$ . The number $b-\delta$ is the magnitude of the jump of $Y_{\delta, b}$ at $b$ and is interpreted as the minimum amount of money that the insured receives if they file a claim. We will refer to contracts with the above form as contracts with constant retention. Clearly, $Y_{\delta, b}$ is not manipulation-proof when $b\gt \delta$ and $Y_{\delta, b}$ is a simple deductible when $b=\delta$ . More importantly, notice that the insured files a reclamation if and only if $Y_{\delta, b}(x)\gt 0$ , so we can partition the interval $[0, M]$ in the two regions $\{Y\gt 0\}$ and $\{Y=0\}=\{Y\gt 0\}^C$ .

Suppose now that the cost $c$ satisfies Assumption2, that the insured cannot use arson-type actions so that there are no incentive compatibility constraints and that we are restricting our search to contracts with constant retention. A few calculations show that we can write the optimization problem as

(Problem FC and No-A) \begin{align} \max _{\delta \geq 0,\, b\geq 0} & \int u(W_0 - H - x +Y_{\delta, b}(x))\,\textrm{d}\mathbb{P} \\[2pt] \nonumber\end{align}
(Positive jump) \begin{align} \textrm{s.t.}\quad & c_0 \mathbb{P}[X\geq b] + \mathbb{E}[X-\delta \vert X\geq b] = (1+\rho )^{-1}H \nonumber \\ & b-\delta \geq 0 . \\[10pt] \nonumber \end{align}

Clearly, the optimization problem with arson-type risks (Problem FC and A) is (Problem FC and No-A) when the Positive jump constraint is binding so $b=\delta$ . It is thus clear that arson-type risks can hurt the insured. We are left to show that arson-type risks always hurt the insured when the fixed cost is strictly positive. This entails showing that the constraint (Positive jump) of (Problem FC and No-A) is an equality if and only if $c_0\gt 0$ . This is a well-known result, being essentially Theorem 2 of Gollier (Reference Gollier1987).

Lemma 5. Consider ( Problem FC and No-A ). It holds that $b=\delta$ if and only if $c_0=0$ .

Proof. If $c_0=0$ then (Problem FC and No-A) is the standard Arrow-Borch-Raviv problem for which the solution is a simple deductible, i.e., $Y(x)=\max \{0, x-d\}$ , so $\delta =b=d$ . Conversely, set $\delta =b=d^*\geq 0$ such that $d^*$ solves the reduced problem

\begin{align*} \max _{d\geq 0} & \int u(W_0 - H -x +\max \{0, x-d\})d\mathbb{P} \\ \textrm{s.t.}\quad & c_0 \mathbb{P}[X\geq d] + \mathbb{E}[X-d\vert X\geq d] = (1+\rho )H . \end{align*}

If $c_0=0$ , then we are done. Suppose $c_0\gt 0$ . Then there exists an $\varepsilon \gt 0$ such that the alternative contract

\begin{equation*}Y^{\prime}(x)=(x-d^*+\varepsilon ){\unicode{x1D7D9}}_{\{x\geq d^*\}}\end{equation*}

strictly improves upon the contract $Y(x)=\max \{0, x-d^*\}$ , as desired.

The interpretation is straightforward. Absent arson-type risks, the optimal insurance contract provides the maximum possible protection while eliminating nuisance claims, the small claims which are more costly to administer than the coverage they offer. This is achieved by refusing to reimburse small claims while providing generous protection conditional on the loss being large. This creates a spread in coverage, the discontinuity at the threshold. But Bob wants to exploit this spread by augmenting the damage to his bike. In equilibrium, Bob is never offered such a contract with high protection, and Bob would be strictly better off if he could commit not to buy a sledgehammer.

Continuous costs and the sub-optimality of disappearing deductibles

We now consider the case when $c$ is a continuous function.

Assumption 3. The cost function $c$ is continuous, weakly positive, non-decreasing, twice-differentiable and satisfies $c(0)=0$ .

Recall that our problem is now almost identical to the one of Spaeter & Roger (Reference Spaeter and Roger1997).

Lemma 6. Under Assumption3, if $Y$ solves the reduced problem

(Reduced problem) \begin{align} \max _{H \geq 0, Y \in B_+(\mathcal{B}([0,M]))} & \int u(W_0 - H - x +Y(x))\,\textrm{d}\mathbb{P} \\ \textrm{s.t.}\, & \,0 \leq Y \leq X \nonumber \\ & (1+\rho )\int Y(x)+c(Y(x)) \,\textrm{d} \mathbb{P} = H \nonumber \end{align}

of Spaeter & Roger (Reference Spaeter and Roger1997) and $Y$ satisfies the slope constraint $\textrm{slope}(Y)\leq (1+\beta )$ (S2) then $Y$ solves (Problem S).

This lemma informs us that the only problematic case that must be handled is when the contract $Y$ found in Spaeter & Roger (Reference Spaeter and Roger1997) does not satisfy constraint (S2) somewhere. Intuitively, it seems natural to attempt flattening $Y$ sufficiently to satisfy $\textrm{slope}(Y)\leq 1+\beta$ , thus solving Problem S. This approach is sometimes fruitful but does not work in some instances, as we explain later. We consider the easiest case when the best contract absent arson-type risks is a completely disappearing deductible. Formally, we say that contract $Y$ is a completely disappearing deductible if there exists a realization $x^{\prime}\in [0, M]$ such that for every $x\geq x^{\prime}$ , it is $Y(x)=x$ .

Proposition 2. If $\beta =0$ , $Y_R$ solves (Reduced problem) and $Y_R$ is a completely disappearing deductible, then the unique solution to (Problem I) is a straight deductible.

Proof. The fact that $Y_R$ is a completely disappearing deductible informs us that our optimal contract $Y$ provides as much coverage as possible while not providing full insurance. Since $\beta =0$ , it holds that any optimal contract $Y$ must have a slope lesser or equal to $1$ . We claim that any optimal contract $Y$ is a simple deductible. This follows from the usual argument that if $Y^{\prime}$ is a generalized deductible with a slope lesser or equal to $1$ , then we can find a deductible contract $Y$ such that $Y\geq _{\textrm{cx}}Y^{\prime}$ and $Y$ (weakly) improves upon $Y^{\prime}$ , where $\geq _{\textrm{cx}}$ denotes the convex order (Shaked & Shanthikumar, Reference Shaked and Shanthikumar2007).

Hence, finding $Y$ is equivalent to solving

\begin{align*} \max _{d\geq 0} & \int u(W_0 - H - x +\max \{0, x-d\})\,\textrm{d} {\mathbb{P}} \\ \textrm{s.t.}\quad &(1+\rho )\mathbb{E}[\max \{0, X-d\} +c(\max \{0, X-d\})] = H, \end{align*}

as desired.

Remark 2. Notice that Proposition 2 is not surprising, as it is essentially the result obtained in Huberman et al. (Reference Huberman, Mayers and Smith1983). Our approach, however, is entirely elementary and does not rely on the Revelation Principle.

As before, Bob’s ability to take a sledgehammer to its bicycles is priced in the insurance contract, and Bob will never be offered a completely disappearing deductible in equilibrium. Again, though Bob would like to purchase such a contract, he cannot. This is because no Bob can commit to being honest. The intuition that the contract-solving problem (Problem S) is simply a “flattening” of the solution to (Reduced problem) is misleading. By additivity of the Lebesgue integral, the insurer’s participation constraint states that the insurer should recoup the cost on average and not realization-by-realization. This, fortunately, gives us some leeway in solving (Problem S). However, this also means that there are some cases where we have to roll up our sleeves and directly attack the problem.

3. On robustness and heuristics

We showed that the optimal contract must be robust to arson-type risks and provided a general characterization of the contract. We now explain why the simplicity of our approach is useful and why we believe both risk theorists and practitioners alike should be interested in it. We focus on risk practitioners, who might want to consider more complex situations than what is envisioned in the text. The main message of the paper is not to worry. The heuristic of considering only contracts mixing deductibles, coinsurance, and upper limits is well-founded when there are arson-type actions, and the risk professionals might want to focus their attention on controlling other risks. A simple example is a car insurer that uses a coinsurance clause to prevent reckless driving (ex-ante moral hazard, see below). As coinsurance contracts are manipulation-proof with regard to arson-type actions, the car insurer should not worry about arson-type risks. It can thus focus on pricing effectively reckless driving. We conclude with a discussion of the limitations of the model that are relevant to the practitioners and with avenues for further research.

Robustness of the mechanism

Readers familiar with mechanism design might wonder why we emphasize that our approach does not invoke Myerson’s Revelation Principle to prove Theorem1. The issue lies in the “robustness” of the result, as the risk practitioner might be interested in situations where the Revelation Principle does not hold. An important example in the literature is the Costly State Falsification framework, where Maggi & Rodriguez-Clare (Reference Maggi and Rodriguez-Clare1995) clarified that the Revelation Principle cannot be applied to “extract” all private information at every stage of the game. Indeed, this fact was extensively used in the subsequent literature on insurance fraud à la Picard (Reference Picard2000). Another example of interest is the literature on games with evidence spawned by Green & Laffont (Reference Green and Laffont1986). In contrast, our elementary argument of manipulation-proofness is simple and easy to adapt to different settings.

Robustness of the decision criteria

A careful inspection of the proof of Theorem1 reveals two important facts. First, notice that we never used any properties of the underlying probability space besides that (i) the random variable representing the risk faced by the insured had full support and that (ii) the two integrals in (Problem I) are finite. In layperson’s terms, we only used the fact that when Bob has an accident, (i) the damage to his bike can fully range from small scratches to a total loss and (ii) that Bob’s risk is not infinitely large. This is important because we could have considered other decision criteria without changing anything to the veracity of Theorem1. For instance, replacing the probability $\mathbb{P}$ with a capacity $\nu$ in one integral and integrating in the sense of Choquet instead of Lebesgue would have changed nothing, provided that (Problem I) still admits a solution and that another technical condition is satisfied.Footnote 10 This is important for risk practitioners because it implies that one does not want to sell insurance contracts vulnerable to arson-type actions, no matter how difficult it is to evaluate the fundamentals of the risk to bear.Footnote 11

Robustness to ex-ante moral hazard in loss reduction

The most well-known and studied agency problem is the classical Principal-Agent problem with hidden actions (ex-ante moral hazard). We refer to Winter (Reference Winter2013) for an introduction to the Principal-Agent problem in the context of insurance design. In insurance, the hidden action is often interpreted as unobservable preventive measures that the insured takes to reduce its risk – driving carefully, for instance. There is no tradeoff in providing incentives to mitigate ex-ante moral hazard and preventing arson-type actions. This is because the optimal contract with only ex-ante moral hazard (often) satisfies the no-sabotage condition and is, therefore, manipulation-proof with regard to arson-type actions. More precisely, the contract satisfies the no-sabotage condition when the first-order approach is valid, a condition often assumed in the literature.

Limitations and avenues for further research

The result that the optimal contract is manipulation-proof seems to be in contradiction with the real world, where arson and insurance fraud do happen. It is not. This apparent contradiction comes from the interplay of three critical assumptions: we defined arson-type actions as (i) the ability to physically destroy an object while (ii) having no chance of being caught and (iii) assuming implicitly that the insured faces no other risks.

While (i) is a natural consideration in the context of Property insurance, it does not capture all the possibilities of insurance fraud. On theoretical grounds, the classical Costly State Falsification model of Crocker & Morgan (Reference Crocker and Morgan1998) contains an intuitive example of where manipulations happen. In their model, the insurer can defraud the contract by declaring greater damages than the real damages, and the insurer cannot verify the claims. The optimal contracts always entail manipulations in equilibrium because the marginal cost of lying is essentially null for small lies. However, the insured’s ability to fill dishonest claims depends on the damage’s size, and the insurer can still differentiate between small and large fundamental losses. This unraveling implies that a contract with manipulations in equilibrium dominates (their model’s) manipulation-proof contracts.Footnote 12

With (ii), we implicitly assumed that arson-type actions cannot be detected by auditing. This is unlikely to be true in general, as arson-type actions often leave pieces of evidence. For instance, fire accelerants like gasoline leave traces that forensic investigators can detect. Given that perfectly preventing arson-type actions by using manipulation-proof contracts is costly in terms of welfare, one might want to control for these actions by auditing suspicious claims instead of implementing manipulation-proof contracts. This opens up an intriguing theoretical possibility where arson-type actions might happen in equilibrium as a result of optimal contracting.

In some situations, arson-type risks do not seem realistic on practical grounds. The leading example is the health insurance field, where arson-type actions, as we modeled them, would require the insured to commit self-injury. However, anecdotes exist where people attempted to defraud an insurance contract by hurting themselves, health insurance contracts routinely contain self-inflicted injury exclusion clauses, and insurance companies routinely investigate injury claims. It is thus not far-fetched to think that self-injuries can be modeled using the method developed in this article. This brings us to (iii), as our manipulation-proof result does not cover the situation where the insured also faces uninsurable (background) risks. This observation also opens up another intriguing theoretical possibility: arson-type actions like self-injuries could arise in equilibrium as a response to significant, uninsurable losses.

Acknowledgements

We thank Nenad Kos, Massimo Marinacci, and Ruodu Wang for their support and Massimiliano Amarante, Mario Ghossoub, David Salib, Richard Peter, and the participants of the 2021 American Risk and Insurance Association’s conference for their comments.

Data availability statement

Data availability is not applicable to this article as no new data were created or analyzed in this study.

Funding statement

Jean-Gabriel Lauzier reports financial support was provided by Bocconi University Department of Economics Ettore Bocconi. Jean-Gabriel Lauzier reports financial support was provided by the University of Waterloo Department of Statistics and Actuarial Science.

Competing interests

Jean-Gabriel Lauzier reports no competing interests to declare.

Footnotes

1 We provide more details on this observation in Section 2.1

2 So $Y\in B_+(\mathcal{B}([0, M]))$ , the space of non-negative and bounded functions (sup-norm) which are measurable with regard to the Borel $\sigma$ algebra $\mathcal{B}$ of $[0, M]$ .

3 The assumption that $z \in [0, M-x]$ implies that the insured cannot generate more losses than the maximum loss $M$ .

4 To be precise, $\sigma _Y$ is a correspondence $[0, M]\ni x\mapsto \sigma _Y(x) \subset [0, M-x]$ .

5 We refer to this type of contract as a contract with constant retention; we will reencounter constant retention contracts in Section 2.2.

6 This follows from Rademacher’s Theorem.

7 The notation $C^0$ denotes the space of continuous functions.

8 The random variables $X_1, X_2$ and $X_3$ are comonotonic if for every $i\neq j$ it is $(X_i(\omega )-X_i(\omega ^{\prime}))(X_j(\omega )-X_j(\omega ^{\prime})) \ge 0$ for $(\mathbb{P}\times \mathbb{P})$ almost every $(\omega, \omega ^{\prime})\in \Omega ^2$ .

9 This statement is true without arson-type risks. We thank an anonymous referee for informing us of this result.

10 The technical condition is that the risk to insure must still have full support under the probabilistic belief of the decision-maker with the non-expected utility criterion. Otherwise, the optimal contract might not be manipulation-proof. This happens when a decision-maker believes that the set of realizations for which a contract induces manipulation is a set of measure zero. Formally, this means we need to require the capacity $\nu$ to be absolutely continuous with regard to $\mathbb{P}$ .

11 Non-expected utility decision criterion can be interpreted as situations where the decision-maker has to infer the risks statistically. This extension does not change the fundamental message of the article; one can keep invoking arson-type actions to justify the no-sabotage condition.

12 The topic of fraud has received a great deal of attention in the literature, and our short discussion is far from exhaustive. We refer to Picard (Reference Picard2013) for an overview.

References

Boonen, T. J., Liu, F., & Wang, R. (2021). Competitive equilibria in a comonotone market. Economic Theory, 72, 12171255.CrossRefGoogle Scholar
Cai, J., & Chi, Y. (2020). Optimal reinsurance designs based on risk measures: A review. Statistical Theory and Related Fields, 4(1), 113.CrossRefGoogle Scholar
Carlier, G., Dana, R.-A., & Galichon, A. (2012). Pareto efficiency for the concave order and multivariate comonotonicity. Journal of Economic Theory, 147(1), 207229.CrossRefGoogle Scholar
Carlier, G., & Dana, R.-A. (2005). Rearrangement inequalities in non-convex insurance models. Journal of Mathematical Economics, 41(4-5), 483503.CrossRefGoogle Scholar
Dana, R.-A. (2003). Pareto efficient insurance contracts when the insurer’s cost function is discontinuous. Economic Theory, 21(4), 871893.Google Scholar
Crocker, K. J., & Morgan, J. (1998). Is honesty the best policy? Curtailing insurance fraud through optimal incentive contracts. Journal of Political Economy, 106(2), 355375.CrossRefGoogle Scholar
Dana, R.-A., & Meilijson, I. (2003). Modelling agents’ preferences in complete markets by second order stochastic dominance, Cahiers du CEREMADE .Google Scholar
Dionne, G. (Ed.). (2000). Handbook of insurance. Boston: Kluwer Academic Publishers, Springer.CrossRefGoogle Scholar
Gollier, C. (1987). Pareto-optimal risk sharing with fixed costs per claim. Scandinavian Actuarial Journal, 1987(1-2), 6273.CrossRefGoogle Scholar
Green, J. R., & Laffont, J.-J. (1986). Partially verifiable information and mechanism design. The Review of Economic Studies, 53(3), 447456.CrossRefGoogle Scholar
Huberman, G., Mayers, D., & Smith, C. W. (1983). Optimal insurance policy indemnity schedules. The Bell Journal of Economics, 14(2), 415426.CrossRefGoogle Scholar
Lacker, J. M., & Weinberg, J. A. (1989). Optimal contracts under costly state falsification. Journal of Political Economy, 97(6), 13451363.CrossRefGoogle Scholar
Landsberger, M., & Meilijson, I. (1994). Co-monotone allocations, Bickel-Lehmann dispersion and the Arrow-Pratt measure of risk aversion. Annals of Operations Research, 52(2), 97106.CrossRefGoogle Scholar
Ludkovski, M., & Rüschendorf, L. (2008). On comonotonicity of Pareto optimal risk sharing. Statistics & Probability Letters, 78(10), 11811188.CrossRefGoogle Scholar
Maggi, G., & Rodriguez-Clare, A. (1995). Costly distortion of information in agency problems. The RAND Journal of Economics, 26(4), 675689.CrossRefGoogle Scholar
Mas-Colell, A., Whinston, M. D., Green, J. R. (1995). Microeconomic theory, vol. 1. Oxford University Press.Google Scholar
Myerson, R. B. (1981). Optimal auction design. Mathematics of Operations Research, 6(1), 5873.CrossRefGoogle Scholar
Picard, P. (2000). On the design of optimal insurance policies under manipulation of audit cost. International Economic Review, 41(4), 10491071.CrossRefGoogle Scholar
Picard, P. (2013). Economic analysis of insurance fraud. In Handbook of insurance (pp. 349395). Springer CrossRefGoogle Scholar
Shaked, M., & Shanthikumar, G. (2007). Stochastic orders. Springer Science & Business Media.CrossRefGoogle Scholar
Spaeter, S., & Roger, P. (1997). The design of optimal insurance contracts: A topological approach. The GENEVA Papers on Risk and Insurance-Theory, 22(1), 519.CrossRefGoogle Scholar
Winter, R. A. (2013). Optimal insurance contracts under moral hazard. In Handbook of insurance (pp. 205230).CrossRefGoogle Scholar
Figure 0

Figure 1 Disappearing v.s. straight deductible.

Figure 1

Figure 2 The value function $V_Y$ of the manipulation stage of the game for $Y(x)=(x - \delta ){\unicode{x1D7D9}}_{\{x \geq b\}}$ for $b\in (0,M)$, $\delta \in (0,b)$ and $\beta \gt 0$.

Figure 2

Figure 3 Straight deductibles and constant retention contracts.