Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
A Bonus-Malus Framework for Cyber Risk Insurance and Optimal Cybersecurity Provisioning Qikun Xiang1 , Ariel Neufeld1 , Gareth W. Peters2 , Ido Nevat3 , and Anwitaman Datta4 arXiv:2102.05568v1 [math.OC] 10 Feb 2021 1 2 Division of Mathematical Sciences, Nanyang Technological University, Singapore Department of Actuarial Mathematics and Statistics, Heriot-Watt University, Edinburgh, UK 3 4 TUMCREATE, Singapore School of Computer Science and Engineering, Nanyang Technological University, Singapore Abstract The cyber risk insurance market is at a nascent stage of its development, even as the magnitude of cyber losses is significant and the rate of cyber risk events is increasing. Existing cyber risk insurance products as well as academic studies have been focusing on classifying cyber risk events and developing models of these events, but little attention has been paid to proposing insurance risk transfer strategies that incentivize mitigation of cyber loss through adjusting the premium of the risk transfer product. To address this important gap, we develop a Bonus-Malus model for cyber risk insurance. Specifically, we propose a mathematical model of cyber risk insurance and cybersecurity provisioning supported with an efficient numerical algorithm based on dynamic programming. Through a numerical experiment, we demonstrate how a properly designed cyber risk insurance contract with a Bonus-Malus system can resolve the issue of moral hazard and benefit the insurer. Keywords—Cyber risk insurance, Cybersecurity, Bonus-Malus, Stochastic optimal control, Dynamic programming 1 Introduction 1.1 The Ever-Increasing Threat of Cyber Crimes Over the years, the frequency and severity of cyber attacks have increased significantly globally, and will continue to increase in the future. Recently, Cybersecurity Ventures estimated the cost of cyber crimes to rise to 10.5 trillion USD annually by 2025 (Morgan, 2020), up from a world economic forum estimate of 3 trillion USD for 2015. The world economic forum’s annual global risk report (Franco, 2020) regularly puts cyber attacks and theft of data in its “Top 5 global risks in terms of likelihood”. Cyber crime is being perpetrated on a massive scale, over a range of different actors in society, hitting individuals in their personal environment as well as organisations. Cyber crime is also a risk type that 1 affects a large array of different organisations worldwide, e.g. government agencies, universities, financial sectors, private corporations, and generally across all industries, including important infrastructure units that play a key role in population security and safety, such as emergency services and health care. Such attacks have caused breaches and significant damages to those organizations, which are vulnerable to intrusion, and often adversely affect downstream users of the compromised services and organizations. The damages incurred can include losses attributed to outcomes such as business interruption, loss of data, reduced reputation and trust of the organisation, legal liabilities, intellectual property theft, and potential for loss of life. These damages result in various degrees of financial loss, including devastating losses as well as ongoing high frequency losses. Cyber attacks can be initiated by both malicious actors within institutions and also external to institutions such as cyber criminals, rogue nation states, hackers, cyber terrorists, and others with malicious intent causing significant negative impact and cost. Cyber attacks come in a variety of forms, ranging from from denial-of-service (DoS) attacks (Gupta and Badve, 2017), malware (Tailor and Patel, 2017), ransomware (Tailor and Patel, 2017), blackmail (Rid and McBurney, 2012), extortion (Young and Yung, 1996), and more (Craigen, Diakun-Thibault, and Purse, 2014; Husák, Komárková, Bou-Harb, and Čeleda, 2019). Many forms of cyber attacks can weaponize third party infrastructure and are not bounded by geographical distance, and hence do not require specialised equipment to devise and initiate. The seriousness of Cyber attacks has been reflected in the U.S. President’s executive order on Strengthening the Cyber security of Federal Networks and Critical Infrastructure, which calls for a cybersecurity framework that can “support the cybersecurity risk management efforts of the owners and operators of the Nation’s critical infrastructure”. Cyber risk from a financial and insurance perspective has also been developed under international banking and insurance regulations, where the Basel III banking accords cover cyber risk as a key component of Operational Risk captial modeling and adequacy, and the Solvency II insurance regulations discuss the significance of an emerging cyber insurance threat that affects both insurers as well as reinsurers. For example, see an overview of cyber risk from a financial and insurance perspective in Peters, Shevchenko, and Cohen (2018b). Financial and governmental regulatory bodies largely classify cyber events according to the following categories: 1. System malfunctions/issue – own system or network is malfunctioning or creating damage to thirdparty’s systems or supplier’s system not functioning, impacting own digital operations; 2. Data confidentiality breach – data stored in own system (managed on premise or hosted/managed by third party) has been stolen and exposed; 3. Data integrity/availability – data stored in own system (managed on premise or hosted/managed by third party) have been corrupted or deleted; 4. Malicious activity – misuse of a digital system to inflict harm (such as cyber bullying over social platforms or phishing attempts to then delete data) or to illicitly gain profit (such as cyber fraud). As an example, consider the Federal Information Security Management Act of 2002 (FISMA), which 2 states their working definition of cyber crime and information security in such a manner as to link the identified operational cybersecurity risks to specific examples of consequences impacting confidentiality, integrity, and availability: “Information Security: means protecting information and information systems from unauthorized access, use, disclosure, disruption, modification, or destruction in order to provide: integrity, which means guarding against improper information modification or destruction, and includes ensuring information non repudiation and authenticity; confidentiality, which means preserving authorized restrictions on access and disclosure, including means for protecting personal privacy and proprietary information; and availability, which means ensuring timely and reliable access to and use of information.” 1.2 The Need for Better Models for Cyber Risk Insurance Although many different security solutions have been developed and implemented in order to detect and prevent cyber attacks, achieving a complete security protection is not feasible (Lu, Niyato, Privault, Jiang, and Wang, 2018b). To address this problem, there is an increasing demand to develop the market for cyber risk insurance and to understand the structuring of insurance products that will facilitate risk transfer strategies in the context of cyber risk and financial risk, see discussions in Peters et al. (2018b); Peters, Shevchenko, and Cohen (2018a); Marotta, Martinelli, Nanni, Orlando, and Yautsiukhin (2017); Böhme and Schwartz (2010) and the references therein. The confluence of increasing sophistication and frequency of cyber attacks on IT infrastructure, the increasing collections of sensitive data in private enterprises and government agencies coupled with the onset of emerging regulatory frameworks, such as Basel II/III and the insurance regulation of Solvency II which contain core requirements related to cyber risk mitigation and modeling have prompted the study of questions pertaining how best to develop a cyber risk insurance market place. As one can see from surveys such as Marotta et al. (2017) and Shetty, Schwartz, Felegyhazi, and Walrand (2010) the scope of such a market place is still very much in its infancy. This has occurred in the insurance space despite the fact that banks and financial institutions rank cyber event losses in their top three loss events systematically when reporting Operational Risk loss events under Basel II/III to national regulators. The reason that the cyber insurance market has yet to emerge with standardised products has largely arisen due to differences in opinion as how best to mitigate and reserve against these loss events. From the IT perspectives, it is common to take a technology perspective to attempt to mitigate such events in contrast to insurance or capital reserving, see discussions in Bandyopadhyay, Mookerjee, and Rao (2009). From a financial risk perspective, risk practitioners see cyber risk from the Operational Risk Basel II/III accord perspective and attempt to avoid insurance mitigation, instead opting for Tier I capital reserving, as the capital reduction from Operational Risk insurance is capped under Basel regulations with a haircut of 20%, see discussions in Peters, Byrnes, and Shevchenko (2011), disincentivizing them to purchase insurance products that have excessive premiums. From the insurance industry’s perspective, there is a lack of market standardisation on insurance contract specifications that would avoid excessive premiums being required to be charged when bespoke insurance products are designed. 3 These three perspectives are beginning to change and we believe it is now a suitable time to revisit the perennial question of how best to set up an insurance market place for cyber risk events. From a technological perspective, cybersecurity is achieved by applying multiple controls which span across being preventive, detective and corrective, and thus realizing defense in depth. For instance, for defense against a distributed denial-of-service (DDoS) attack which would violate the availability property of a service, an organization may opt to use network traffic filtering as well as a content distribution network, and use multiple servers to balance the network load. The life-cycle of a data breach event often involves the initial intrusion event, subsequent escalation of privileges till the exfiltration of data. The victim thus has a window of opportunity from the moment of intrusion to the eventual breach of the data, to detect and prevent it. An organization may also invest in storing the data encrypted. As such, there are various controls for ensuring confidentiality property in this instance, but each control adds to the upfront costs of risk prevention or reduction. An organization needs to determine the quantum of its security (risk prevention) budget and distribute this across these controls. We argue that this budget for risk reduction needs to and can be determined in tandem with risk transfer decisions. In particular, we want to explore a model where an organization enjoys a reduction in the cost of risk transfer as a result of its upfront expenses in risk reductions; and like-wise an insurer benefits from a pricing model which encourages good security posture among its customers. 1.3 The Proposed Cyber Risk Framework We propose in this paper a perspective that combines classical market place structuring via a BonusMalus framework that provides IT-specific incentive mechanisms that act to encourage sound IT governance and technology developments, whilst also allowing insurers to encourage risk reduction in their risk pools to provide competitive insurance premium pricing. We introduce for the first time a cyber riskbased Bonus-Malus framework and then demonstrate how to develop loss models and decision making models under uncertainty in this framework from both individual and insurance providers’ perspectives. 1.4 Related Work Recently, many studies analyzed the cyber risk insurance from a technology perspective. Security frameworks involving cyber risk insurance have been developed for specific IT systems, including computer networks (Fahrenwaldt, Weber, and Weske, 2018; Xu and Hua, 2019), heterogeneous wireless network (Lu, Niyato, Jiang, Wang, and Poor, 2018a), wireless cellular network (Lu et al., 2018b), plug-in electric vehicles (Hoang, Wang, Niyato, and Hossain, 2017), cloud computing (Chase, Niyato, Wang, Chaisiri, and Ko, 2019), and fog computing (Feng, Xiong, Niyato, Wang, and Leshem, 2018). Some studies considered the interplay between self-mitigation measures (i.e. risk reduction) and cyber risk insurance, e.g. Pal and Golubchik (2010); Pal, Golubchik, Psounis, and Hui (2014, 2019); Khalili, Naghizadeh, and Liu (2018); Dou, Tang, Wu, Qi, Xu, Zhang, and Hu (2020). These studies investigated two important challenges in cyber risk insurance: risk interdependence and moral hazard. They found that in order to incentivize the insured to invest in self-mitigation measures, some form of contract 4 discrimination, i.e. adjusting the insurance premium based on the insured’s security investment, is necessary. Yang and Lui (2014); Schwartz and Sastry (2014); Zhang, Zhu, and Hayel (2017) investigated these challenges in a networked environment, where cyber attacks can spread between neighboring nodes, further complicating these challenges. There are also studies which took the insured’s perspective and analyzed the security provisioning process using dynamic models. Chase et al. (2019) developed a framework based on stochastic optimization to jointly provision cyber risk insurance and cloud-based security services across multiple time periods in cloud computing applications. Zhang and Zhu (2018) modeled the decisions on self-protections of the insured by a Markov decision process and investigated the problem of insurance contract design. A critical drawback of many of the existing studies is that they neglected the highly uncertain nature of losses incurred by cyber incidents. These studies relied on over-simplified assumptions, e.g. by modeling cyber loss as: a fixed amount (Pal and Golubchik, 2010; Pal et al., 2014; Yang and Lui, 2014; Hoang et al., 2017; Feng et al., 2018; Dou et al., 2020), random with finite support (Chase et al., 2019; Zhang and Zhu, 2018; Lu et al., 2018b), or random with a simple parametric distribution (Zhang et al., 2017; Khalili et al., 2018). These assumptions limit the practicality of these studies, and their results remain conceptual and non-applicable to realistic insurance loss modeling under a classical Loss Distributional Approach (LDA) framework. Moreover, many of these studies do not take into account the interplay between the upfront costs of prevention, the consequent reduced risks, and the possibility to exploit this interplay to design practical cyber risk insurance products. A review of cyber insurance product prospectus by major insurers, such as AIG’s CyberEgde1 , Allianz Cyber Protect2 , and Chubb’s Cyber Enterprise Risk Management (Cyber ERM)3 indicates that the insurance products in the market have yet to explicitly factor in the benefits of up-front protection, or to incentivize and offset those costs against that of risk transfer. We introduce the Bonus-Malus system which is frequently used in vehicle insurance products to address this gap in cyber risk product design. 1.5 Contributions Our main contributions are as follows: 1. We introduce the Bonus-Malus system to cyber risk insurance as a mechanism to provide incentive for the insured to adopt self-mitigation measures against cyber risk. 2. We develop a mathematical model of cyber losses and cyber risk insurance, and subsequently analyze the optimal cybersecurity provisioning process of the insured under the stochastic optimal control framework. 3. We develop an efficient algorithm based on dynamic programming to accurately solve the stochastic optimal control problem, under the assumption that the loss severity follows a truncated version 1 https://www.aig.com/business/insurance/cyber-insurance/, accessed on 2020-12-10 2 https://www.agcs.allianz.com/solutions/financial-lines-insurance/cyber-insurance.html, accessed on 2020- 12-10 3 https://www.chubb.com/us-en/business-insurance/cyber-enterprise-risk-management-cyber-erm.html, accessed on 2020-12-10 5 of the g-and-h distribution. We also formally prove the correctness of the proposed algorithm. 4. We demonstrate through a numerical experiment that a properly designed cyber risk insurance contract with a Bonus-Malus system can resolve the issue of moral hazard, and can provide benefits for the insurer. 1.6 Organization of the Paper The rest of the paper is organized as follows. In Section 2, we introduce the mathematical model of cyber losses and cyber risk insurance with a Bonus-Malus system. In Section 3, we present the optimal cybersecurity provisioning process and the dynamic programming algorithm. In Section 4, we introduce the g-and-h distribution and use it as the model for loss severity. We present results from the numerical experiment in Section 5. Finally, Section 6 concludes the paper. 2 Cyber Risk Insurance Policy and Bonus-Malus System Let us first present an overview of our cyber risk insurance model, which begins with a specification of the frequency and severity model under consideration in the Loss Distributional Approach (LDA) that defines the financial loss process that results from cyber risk events. We consider T ∈ N consecutive years, and we assume that throughout each year t, the insured may suffer a random number (Nt ∈ N) (t) (t) of cyber loss events arising from cyber attack events. These loss amounts are denoted by X1 , . . . , XNt . PNt (t) Xk . Such a loss model is referred to as the The cumulative annual cyber loss in year t is therefore k=1 Loss Distributional Approach (LDA) in the study of Operational Risk. The insured has several choices to attempt to mitigate these cyber events and reduce the risk, including enhancing the security and resilience of their IT infrastructure and reserving Tier I capital to cover the incurred losses. In regards to the internal IT infrastructure, it will be assumed that the insured has the option to adopt a self-mitigation measure that can reduce the severity of cyber incidents up to a fixed amount of loss corresponding to the effect of the self-mitigation measure. In addition, the insured can choose to purchase a cyber risk insurance policy which gives the insured the right to claim the cumulative cyber loss incurred, up to a maximum cap imposed by the insurance contract, in an agreed interval of time (typically annually), minus a deductible. 2.1 Cyber Loss Model Let us define W := S n n∈Z+ {n} × R+ to be the space representing all possible combinations of realizations of the number of events per year (frequency) and individual losses per event (severities). Let B(W) := S σ( n∈Z+ {(n, B) : B ∈ B(Rn+ )}) be a σ-algebra on W, where B(Rn+ ) denotes the Borel subsets of Rn+ . (1) (1) Let us consider the space Ω := (W)T = W × · · · × W . For each ω = (w1 , . . . , wT ) = ((n1 , x1 , . . . , xn1 ), {z } | T times (T ) (T ) . . . , (nT , x1 , . . . , xnT )) ∈ Ω, we define Wt (ω) := wt and Nt (ω) := nt for t = 1, . . . , T . Let P1 be a probability measure on (W, B(W)), where the subscript “1” indicates that it is a probability measure 6 for the cyber incidents occuring in a single year. Let P := P1 ⊗ · · · ⊗ P1 , and let (Ft )t=0:T be a filtration | {z } T times on Ω, defined by F0 := {∅, Ω}, Ft := σ((Ws )s=1:t ). Then, (Ω, FT , P, (Ft )t=0:T ) is a filtered probability space. Under these definitions, W1 , . . . , WT are independently and identically distributed (i.i.d.) random variables. Let ψN (s) := E[sN1 ] denote the probability generating function (pgf) of the loss frequency distribution. We assume that for t = 1, . . . , T , (1) (T ) (t) (T ) P[{ω = ((n1 , x1 , . . . , x(1) n1 ), . . . , (nT , x1 , . . . , xnT )) : nt = n, xk ≤ zk , k = 1, . . . , n}] =P[Nt = n] n Y (1) FX (zk ), k=1 where FX (·) is a distribution function corresponding to the severity distribution. This implies that given the loss frequency of a year, the individual loss amounts in that year are i.i.d. We assume R that the severity distribution has finite expectation, i.e. R |x|dFX < ∞. For convenience, we write (t) (t)  Wt = Nt , X1 , . . . , XNt . We use W to refer to a random variable that has the same distribution as W1 , . . . , WT . Similarly, we use N to refer to a random variable that has the same distribution as N1 , . . . , NT , and we use X to refer to a random variable that has distribution function FX . 2.2 Self-Mitigation Measures Let us assume that there exists D ∈ N different self-mitigation measures, and the insured makes the decision to either adopt one of the self-mitigation measures or to not adopt any self-mitigation measure at the beginning of each year. The self-mitigation measure d ∈ D := {0, 1, . . . , D} requires an annual investment of β(d) ∈ R+ per year, and decreases the severity of each cyber loss by up to γ(d) ∈ R+ , that is, the severity of a loss will be decreased from X to (X − γ(d)) +4 with the adoption of the self-mitigation measure d. We assume that β(0) = γ(0) = 0. Thus, if the insured decides to adopt the self-mitigation measure d in a year, then the total loss suffered by the insured that year is given by L(d, w) := n X k=1 + (xk − γ(d)) , (2) when the corresponding loss frequency and severity that year is w = (n, x1 , . . . , xn ) ∈ W. 2.3 Cyber Risk Insurance Policy Let us now consider a cyber risk insurance contract that lasts for T years. At the beginning of each year, the insured decides whether to activate the contract. In the case that the contract has been activated in a previous year, this corresponds to the insured deciding whether to continue the contract. If the contract is activated, the insured pays a premium to the insurer at the start of the year, in exchange for the insurance coverage throughout the year. If the insured decides to withdraw from the contract, it no longer pays the premium and the contract is deactivated so that the insured receives no coverage. We further assume that the insured pays the insurer an initial sign-on fee for fixed costs and contract origination the first time a contract is initiated, in addition to the premium, and that this amount varies 4 Throughout the paper, we use the following notations: (x)+ := max{x, 0}, x ∨ y := max{x, y} and x ∧ y := min{x, y}. 7 deterministically over time and will be denoted by δin (t) ≥ 0 in year t. This can be used to incentivize the insured to activate the contract early. Furthermore, we also assume that the insured pays the insurer a deterministic and time-dependent penalty, denoted by δout (t) ≥ 0, when it withdraws from the contract in year t. Once withdrawn, the insured may re-activate the contract at a later year with a fixed penalty δre ≥ 0. Suppose that the cumulative cyber loss suffered by the insured in a year is L. At the end of the year, the insured decides whether to make a claim to the insurer. Once the claim is processed, the insured receives a payment of (L − ldtb )+ ∧ lmax from the insurer as compensation, that is, the insured covers the loss up to the deductible ldtb ≥ 0, and the insurer covers all of the remaining loss up to the maximum compensation lmax ≥ 0. 2.4 Bonus-Malus System We now introduce the Bonus-Malus system to cyber risk insurance markets. Let us assume that there are B Bonus-Malus levels in the contract, denoted by B := {−B, . . . , −1, 0, 1, . . . , B}, where B + B + 1 = B. At t = 0, the insured starts in the initial Bonus-Malus level, denoted by b0 = 0. At the end of the t-th year, given that the contract is still active, the insurer determines the Bonus-Malus level of the insured based on its previous level bt−1 , and the amount of insurance claims Ct that was given out to this insured in the t-th year, that is, bt = BM(bt−1 , Ct ), where BM : B ×R+ → B denotes the deterministic rules that are transparent to the insured at the signing of the contract. We make the assumption that BM(b, C) is non-decreasing in C for each b ∈ B. Even when the insured has withdrawn from the contract, we assume that the Bonus-Malus level is still updated annually. Concretely, let us define I := {no, on, off1 , . . . , offT } as the set of all possible states of the cyber risk insurance contract. In I, “no” denotes that the contract has not been signed yet, “on” denotes that the contract is continued, and “offy ” denotes that the contract is withdrawn and y ∈ Z+ represents the counter variable that is updated annually as long as the insured does not re-activate the contract. Let BM0 : B × I → B × I be a deterministic transition function that represents the update rules after the insured withdraws from the contract. At the end of the t-th year, given that the contract is inactive, the insurer determines the Bonus-Malus level bt and the insurance state it of the insured based on its Bonus-Malus level and the insurance state in the previous year, that is, (bt , it ) = BM0 (bt−1 , it−1 ). Since no such update is possible before the insured activates the contract, it is required that BM0 (b, no) = (b, no) for all b ∈ B. With the addition of the Bonus-Malus system to the cyber risk insurance, we assume that the premium depends on both time and the current Bonus-Malus level of the insured, and is given by pBM (b, t), where pBM : B × {1, . . . , T } → R+ is a deterministic function. The deductible and the maximum compensation are also assumed to be dependent on both time and the Bonus-Malus level, and are given BM BM by deterministic functions ldtb : B × {1, . . . , T } → R+ and lmax : B × {1, . . . , T } → R+ . We define the function λBM : B × {1, . . . , T } × R+ → R+ by BM BM λBM (b, t, l) := (l − ldtb (b, t))+ ∧ lmax (b, t) to simplify the notation when modeling the loss covered by the insurer. 8 (3) 3 Optimal Cybersecurity Provisioning and Stochastic Optimal Control 3.1 Cybersecurity Provisioning Process Now, having introduced the model for cyber losses and cyber risk insurance, we consider the problem of optimal cybersecurity provisioning from the insured’s point of view. It is assumed that the cybersecurity provisioning process takes place for T consecutive years (same as the length of the cyber risk insurance), and each year consists of the three following stages: 1. Provision Stage. The insured decides: (i) the self-mitigation measure to adopt in this year, denoted by dt ∈ D and (ii) whether to activate/withdraw/rejoin the cyber risk insurance contract, denoted by ιt ∈ {0, 1}. 2. Operation Stage. The random cyber risk events and the corresponding cyber losses suffered by the insured in this year, denoted by Wt , is realized in this stage according to the model described in Section 2.1 and Section 2.2. 3. Claim Stage. If the insurance contract is active, the insured decides whether or not to make a claim, denoted by jt ∈ {0, 1}. In the case where a claim is made, the insured receives a compensation from the insurer corresponding to the cyber loss that year. Subsequently, the Bonus-Malus level of the insured is updated. Formally, let n Π := π = (dt , ιt ,jt )t=1:T : dt : Ω → D, ιt : Ω → {0, 1} are Ft−1 -measurable, jt : Ω → {0, 1} is Ft -measurable, {ιt = 0, jt = 1} = ∅, for t = 1, . . . , T o (4) denote the set of all possible decision policies. The conditions in the above definition are explained as follows: • The decisions dt , ιt are made before observing Wt , hence may depend on all available information up to year t − 1; • The decision jt is made after observing Wt , hence may depend on all available information up to year t; • The condition {ιt = 0, jt = 1} = ∅ requires that the insured may only make a claim (i.e. jt = 1) when the insurance is adopted (i.e. ιt = 1) in year t. For t = 1, . . . , T , let ft : B × I × D × {0, 1} × {0, 1} × W → B × I be the state transition function for each year t, given by ft (b, i, d, ι, j, w) :=     BM(b, jλBM (b, t, L(d, w))), on ,  BM0 (b, i) 9 if ι = 1, if ι = 0. (5) For t = 1, . . . , T , let gt : B × I × D × {0, 1} × {0, 1} × W → R+ be the cost function for each year t, given by gt (b, i, d, ι, j, w) :=β(d) + ιpBM (b, t) + δin (t)✶{i=no,ι=1} + δout (t)✶{i=on,ι=0} + δre ✶{i6=on,i6=no,ι=1} + L(d, w) − ιjλBM (b, t, L(d, w)), (6) where β(d) is the investment of adopting the self-mitigation measure d, ιpBM (b) corresponds to the cyber risk insurance premium, δin (t)✶{i=no,ι=1} + δout (t)✶{i=on,ι=0} + δre ✶{i6=on,i6=no,ι=1} corresponds to the entrance/withdrawal costs, L(d, w) is the total cyber loss, and ιjλBM (b, t, L(d, w)) corresponds to the compensation from the insurer. Now, for any decision policy π = (dt , ιt , jt )t=1:T ∈ Π, let us define the (Ft )t=0:T -adapted controlled  stochastic process bπt , iπt t=0:T as follows: (bπ0 , iπ0 ) :=(0, no), (7) (bπt , iπt ) :=ft (bπt−1 , iπt−1 , dt , ιt , jt , Wt ) for t = 1, . . . , T. Then, gt (bπt−1 , iπt−1 , dt , ιt , jt , Wt ) corresponds to the cybersecurity cost in year t. For π ∈ Π, and t = 1, . . . , T , define Vtπ by Vtπ := E hP T −(s−t)r gs (bπs−1 , iπs−1 , ds , ιs , js , Ws ) s=t+1 e i Ft , (8) where 0 < e−r ≤ 1 is the discount factor and VTπ := 0. We assume that in the cybersecurity provisioning process, the objective of the insured is to minimize the expected value of the discounted total cybersecurity cost. This is formulated as the following finite horizon stochastic optimal control problem: V0 := inf V0π = inf E π∈Π 3.2 π∈Π hP T −tr gt (bπt−1 , iπt−1 , dt , ιt , jt , Wt ) t=1 e i . (9) Dynamic Programming Algorithm The stochastic optimal control problem introduced in Section 3.1 can be solved efficiently via the dynamic programming algorithm. The dynamic programming algorithm iteratively solves a sequence of onestage optimization problems and also constructs an optimal decision policy from the optimizers of these problems. In the following, we define the values of these one-stage optimization problems, denoted by (Vt )t=0:T , and their corresponding optimizers (dbt , b ιt , b jt )t=1:T . For t = T, T − 1, . . . , 0, let Vt : B × I → R+ be recursively defined as follows: for every b ∈ B, i ∈ I, let VT (b, i) :=0, Vt−1 (b, i) :=e−r min d∈D,ι∈{0,1}    n o E min gt (b, i, d, ι, j, W ) + Vt ft (b, i, d, ι, j, W ) . (10) j∈{0,1} For t = 1, . . . , T , let dbt : B × I → D, b ιt : B × I → {0, 1}, and b jt : B × I × W → {0, 1} be defined as 10 follows: for every b ∈ B, i ∈ I, let    n  o b dt (b, i), b ιt (b, i) ∈ arg min E min gt (b, i, d, ι, j, W ) + Vt ft (b, i, d, ι, j, W ) , (11) j∈{0,1} d∈D,ι∈{0,1}  h i   1 if b ιt (b, i) = 1, gt (b, i, dbt (b, i), 1, 1, w) + Vt ft (b, i, dbt (b, i), 1, 1, w)    h i b bt (b, i), 1, 0, w) + Vt ft (b, i, dbt (b, i), 1, 0, w) , jt (b, i, w) := < g (b, i, d t     0 otherwise. (12) The following theorem shows the construction of an optimal decision policy by dynamic programming. Theorem 3.1. Let (Vt )t=0:T be defined as in (10), let the functions (dbt , b ιt , b jt )t=1:T be defined as in (11) and (12), and let π ⋆ = (d⋆t , ι⋆t , jt⋆ )t=1:T ∈ Π be recursively defined as follows: ⋆ bπ0 , iπ0 ⋆  := (0, no), (13) for t =1, . . . , T, let: ⋆ ⋆  ⋆ d⋆t := dbt bπt−1 , iπt−1 , Then, the following holds:  ⋆ ⋆ ⋆  jt⋆ := b jt bπt−1 , iπt−1 , Wt . ι⋆t := b ιt bπt−1 , iπt−1 , ⋆ V0 (0, no) = V0π = V0 . (14) Proof. See Appendix A. As a consequence of Theorem 3.1, let us now introduce an algorithm based on the dynamic programming principle to solve the stochastic optimal control problem (9), which is presented in Algorithm 1. In addition to an optimal decision policy π ⋆ , Algorithm 1 also outputs the state transition probabilities and ⋆ ⋆ marginal state occupancy probabilities of the Markov process bπt , iπt t=0:T , as well as other quantities of interest (see Theorem 3.2(iii)). Theorem 3.2 shows the correctness of Algorithm 1, whereas Remark 3.4 shows its computational tractability. Theorem 3.2. Let V0 (0, no), π ⋆ , Pt⋆  t=1:T ⋆ , Pt t=0:T (m)  , ζt output of Algorithm 1. Then, the following statements hold: t=1:T,m=1:M , and Z ζ (m)  m=1:M be the ⋆ (i) V(0, no) = V0 and π ⋆ is an optimal decision policy, i.e. V(0, no) = V0π = V0 . ⋆ (ii) bπt , iπt ⋆  t=0:T is a discrete-time time-inhomogeneous Markov chain with transition kernels Pt⋆ that is, for all t = 1, . . . , T , (b, i), (b′ , i′ ) ∈ B × I, it holds that  t=1:T i h ⋆ ⋆   ⋆ ⋆  Pt⋆ (b, i) → (b′ , i′ ) = P bπt , iπt = (b′ , i′ ) bπt−1 , iπt−1 = (b, i) . (15) i h ⋆ ⋆ ⋆ P t (b, i) = P bπt , iπt = (b, i) . (16) Moreover, for all t = 0, . . . , T , (b, i) ∈ B × I, it holds that 11 ,  (m)  (iii) Assume that for all m = 1, . . . , M , t = 1, . . . , T , b ∈ B, i ∈ I, it holds that E ζt (b, i, W ) < ∞. Then, for all m = 1, . . . , M , t = 1, . . . , T , it holds that (m) ζt   (m) π⋆ π⋆ =E ζt bt−1 , it−1 , Wt , Z ζ (m) = T X (m) ζt . (17) (18) t=1 Proof. See Appendix A. Example 3.3 (Quantities of interest). Below is a list of quantities that satisfy the assumption of Theorem 3.2(iii). (i) For d ∈ D, let ζt (b, i, w) =   ⋆  ⋆ ✶{dbt (b,i)=d} . Then, we have ζ t = P dbt bπt−1 , iπt−1 = d , which cor- responds to the probability that the self-mitigation measure d ∈ D is adopted in year t under the decision policy π ⋆ . (ii) Let ζt (b, i, w) = e−tr β(dbt (b, i)), then Z ζ corresponds to the expected value of the discounted total self-mitigation investment.    (iii) Let ζt (b, i, w) = e−tr b ιt (b, i) pBM (b) + δin (t)✶{i=no} + δre ✶{i6=no,i6=on} + (1 −b ιt (b, i))δout (t)✶{i=on} , then Z ζ corresponds to the expected value of the discounted total payment from the insured to the insurer. (iv) Let ζt (b, i, w = (n, x1 , . . . , xn )) = e−tr hP n k=1 i xk − L(dbt (b, i), w) , then Z ζ corresponds to the ex- pected value of the discounted total loss that is prevented by adopting self-mitigation measures. i h jt (b, i, w)λBM (b, t, L(dbt (b, i), w)) , then Z ζ corresponds to the expected value (v) Let ζt (b, i, w) = e−tr b of the discounted total insurance compensation the insured receives. Remark 3.4. Assume that the following quantities either admit an analytically tractable expression, or can be efficiently approximated to high numerical precision:     (i) the expectation E L(d, W ) = E[N ]E (X − d)+ , where d ∈ D;   +  (ii) the expectation E ✶I (λBM (b, t, L(d, W ))) λBM (b, t, L(d, W )) − α , where b ∈ B, d ∈ D, t ∈ {1, . . . , T }, α ≥ 0, and I ⊂ R+ is an interval;   (iii) the probability P λBM (b, t, L(d, W )) ∈ I , where b ∈ B, d ∈ D, t ∈ {1, . . . , T }, and I ⊂ R+ is an interval;  (m)  (iv) the expectation E ζt (b, i, W ) , where m ∈ {1, . . . , M }, t ∈ {1, . . . , T }. Then, Algorithm 1 is computationally tractable, meaning that quantities in Algorithm 1 can either be computed exactly or efficiently approximated to high numerical precision. In addition, since λBM (b, t, L(d, W )) BM is bounded above by lmax (b, t), we can assume without loss of generality that the interval I in (ii) and (iii) above is bounded. A concrete model in which Algorithm 1 is computationally tractable will be introduced in Section 4. 12 Algorithm 1: Dynamic Programming for Optimal Cybersecurity Provisioning  BM BM Input: B, I, D, BM, BM0 , pBM , ldtb , lmax , β(·), γ(·), r, ζ (m) (·, ·, ·) m=1:M   (m)  ⋆ , Z ζ (m) m=1:M Output: V0 (0, no), π ⋆ , Pt⋆ t=1:T , P t t=0:T , ζ t t=1:T,m=1:M 1 VT (b, i) ← 0 for all b ∈ B, i ∈ I. 2 for t = T, T − 1, . . . , 1 do for b ∈ B do 3 b ← BM(b, 0), b ← max{BM(b, c) : c ∈ R+ }. 4 for b ≤ b′ ≤ b do 5 αt (b, b′ ) ← Vt (b′ , on) − Vt (b, on), Lt (b, b′ ) ← {c ∈ R+ : BM(b, c) = b′ , c > αt (b, b′ )}. 6 for i ∈ I do 7 for d ∈ D do 8 Ht (b, i, d, 1) ← Vt (b, on) −   +  P BM ′ BM ′ E ✶ λ (b, t, L(d, W )) − α (b, b ) . t {BM(b,λ (b,t,L(d,W )))=b } b≤b′ ≤b 9  Ht (b, i, d, 0) ← Vt BM0 (b, i) . 10  (dbt (b, i), b ιt (b, i)) ← arg mind∈D,ι∈{0,1} β(d) + ιpBM (b, t) + δin (t)✶{i=no,ι=1} +    δout (t)✶{i=on,ι=0} + δre ✶{i6=on,i6=no,ι=1} + E L(d, W ) + Ht (b, i, d, ι) .  −r Vt−1 (b, i) ← e mind∈D,ι∈{0,1} β(d) + ιpBM (b, t) + δin (t)✶{i=no,ι=1} +    δout (t)✶{i=on,ι=0} + δre ✶{i6=on,i6=no,ι=1} + E L(d, W ) + Ht (b, i, d, ι) .  b jt (b, i, w) ← ✶{bιt (b,i)=1} ✶Sb≤b′ ≤b Lt (b,b′ ) λBM (b, t, L(dbt (b, i), w)) .   Pt⋆ (b, i) → (b′ , on) ← 0 for all (b′ , i′ ) ∈ B × I. 11 12 13 14 if b ιt (b, i) = 1 then 15 for b < b′ ≤ b do h i   Pt⋆ (b, i) → (b′ , on) ← P λBM (b, t, L(dbt (b, i), W ))) ∈ Lt (b, b′ ) .     P Pt⋆ (b, i) → (b, on) ← 1 − b<b′ ≤b Pt⋆ (b, i) → (b′ , on) . 16 17 18 else 19   Pt⋆ (b, i) → BM0 (b, i) ← 1. 20 ⋆ ⋆ 21 P 0 (0, no) ← 1, P 0 (b, i) ← 0 for all (b, i) 6= (0, no). 22 for t = 1, 2, . . . , T do 23 24 25 26 27 P ⋆ ⋆ For all (b, i) ∈ B × I, P t (b, i) ← (b′ ,i′ )∈B×I Pt⋆ [(b′ , i′ ) → (b, i)]P t−1 (b′ , i′ ). P (m) ⋆ (m) For m = 1, . . . , M , ζ t ← (b,i)∈B×I E[ζt (b, i, W )]P t−1 (b, i). PT (m) For m = 1, . . . , M , Z ζ (m) ← t=1 ζ t . Define π ⋆ = (d⋆t , ι⋆t , jt⋆ )t=1:T as in (13).   (m)  ⋆ return V0 (0, no), π ⋆ , Pt⋆ t=1:T , P t t=0:T , ζ t , Z ζ (m) m=1:M . t=1:T,m=1:M 13 3.3 Discussion About the Pricing of the Insurance Premium One important consideration of the cyber insurer is the choice of the annual premium. Normally, for a fixed self-mitigation measure d ∈ D, a fixed deductible ldtb ≥ 0, and a fixed maximum compensation lmax ≥ 0, one may consider the risk premium E[(L(d, W ) − ldtb )+ ∧ lmax ]. However, these quantities are variable in our dynamic model with the Bonus-Malus system. Moreover, the insured may choose to change the self-mitigation measure adopted in each year, or withdraw from the cyber risk insurance contract, thus further complicating the matter. From the perspective of the insurer, one option is to set the premium such that the discounted expectation of the difference between the total payment from the insured to the insurer and the total insurance compensation is maximized. For simplicity, let Z ins be defined as the value of Z ζ in Example 3.3(iii) and let Z cp be defined as the value of Z ζ in Example 3.3(v). Hence, Z ins − Z cp corresponds to the discounted expectation of the difference between the total payment from the insured to the insurer and the total insurance compensation, i.e. the insurer’s expected discounted profit, when the insured acts optimally. Notice that Z ins − Z cp ≤ 0 due to the assumption that the insured minimizes the expected value of the discounted total cybersecurity cost. In particular, in the absence of any regulatory mandated requirements or external requirements, a rational insured will only consider adopting the cyber risk insurance if the expected value of the discounted benefit outweighs the expected value of the discounted cost. Therefore, when the premium is set too high, the insured will choose not to adopt the cyber risk insurance and hence Z ins = Z cp = 0. This is clearly undesirable. Another important consideration of the insurer is the retention of customers, since the insurer needs a large homogeneous pool of risk to function. When the premium is high, the insured may withdraw from the contract early due to a transition into a higher Bonus-Malus level. This issue, however, can be addressed by imposing a large withdrawal penalty δout (t), especially for later policy years (i.e. when t is close to T ). We will discuss about the problem of setting the premium with a concrete example in Section 5. We would like to remark that Z ins − Z cp ≤ 0 does not imply that cyber risk insurance is impractical. These quantities are derived under the assumption that the insured follows the optimal cybersecurity provisioning policy. In reality, the insured will typically act sub-optimally either due to the practical complexity of the required computation being beyond the scope of decision makers or due to impartial or incomplete information as to the severity of the risk they face. Insurers offering such products will however build up a loss database which gives them a competitive advantage in knowing the true severity and frequency of such events, for their potential customer base. Regulatory transparency requirements will also play an important role in determining the profit margins that may arise if such products are issued. 4 Modeling Cyber Loss with Truncated g-and-h Distribution In this section, we adopt specific distributional assumptions about the random variable X, which corresponds to the severity of a single cyber risk event. Studies such as Maillart and Sornette (2010) and Wheatley, Maillart, and Sornette (2016) have shown that the severity of cyber risk events have 14 heavy-tailed distributions. It is well-known that the g-and-h family introduced by Tukey (1977) contains distributions with a wide range of skewness and kurtosis (e.g. see Figure 3 of Dutta and Perry (2006)), which makes it suitable for modeling Operational Risk (Dutta and Perry, 2006; Peters and Sisson, 2006; Cruz, Peters, and Shevchenko, 2015; Peters and Shevchenko, 2015). In Dutta and Perry (2006), the following advantages of the g-and-h distribution are discussed: • it is flexible and it fits well to real data under many different circumstances, e.g. when considering losses of the company as a whole and when considering individual business lines or event types; • it produces realistic estimations of the Operational Risk capital; • it is easy to simulate random samples from. The parameters in the g-and-h distribution can be robustly estimated based on quantiles (Xu, Iglewicz, and Chervoneva, 2014) or L-moments (Peters, Chen, and Gerlach, 2016). Due to these properties, we adopt the g-and-h distribution as a particular model for the severity of cyber risk events in this section. The g-and-h distribution is a four-parameter family of distributions, given by the following definition: e follows a g-and-h(α, ς, g, h) distribution, if X e =α + ςYg,h (Z), X where Z ∼Normal(0, 1),      exp(gz)−1 exp hz2 g 2 Yg,h (z) :=    z exp hz2 2 (19) if g 6= 0, if g = 0, where α ∈ R is the location parameter, ς > 0 is the scale parameter, g ∈ R is the skewness parameter, and h ≥ 0 is the kurtosis parameter. In this paper, we assume that the parameters α, ς, g, and h are e is given by fixed and known. By (19), the distribution function of X   e ≤ x] = Φ Y −1 x−α , FXe (x) :=P[X g,h ς (20) −1 where Yg,h denotes the inverse function of Yg,h , and Φ denotes the distribution function of the standard −1 normal distribution. Even though Yg,h cannot be expressed analytically, it can be efficiently evaluated using a standard root-finding procedure such as the bisection method and the Newton’s method. There−1 fore, we treat Yg,h as a tractable function. The g-and-h distribution has the property that the m-th e exists when h < moment of X 1 m (e.g. see Appendix D of Dutta and Perry (2006)). Since we consider losses that are positively skewed and have finite expectation, from now on, we assume that g > 0 and 0 ≤ h < 1. Since cyber losses are positive, we introduce a truncated version of the g-and-h distribution. Definition 4.1 (Truncated g-and-h distribution). For α ∈ R, ς > 0, g > 0, h ∈ [0, 1), the random variable X has truncated g-and-h distribution with parameters α, ς, g, h, denoted by X ∼ Tr-g-and-h(α, ς, g, h), if X has distribution function e ∼ g-and-h(α, ς, g, h). where X e ≤ x|X e > 0], FX (x) := P[X ≤ x] = P[X 15 (21) The next lemma shows some useful properties of the truncated g-and-h distribution. Lemma 4.2. Suppose that X ∼ Tr-g-and-h(α, ς, g, h) for α ∈ R, ς > 0, g > 0, h ∈ [0, 1). Then, the following statements hold. (i) The distribution function of X is given by    FXf(x)−FXf(0) 1−FX f(0) FX (x) =  0 if x > 0, (22) if x ≤ 0, where FXe is defined in (19). (ii) Suppose that U ∼ Uniform[0, 1], and let    XU := α + ςYg,h Φ−1 U + (1 − U )FXe (0) , (23) then XU ∼ Tr-g-and-h(α, ς, g, h).   (iii) For any γ ≥ 0, the expectation E (X − γ)+ is given by: "      g2 g ς −1 + √ exp Φ − Yg,h E (X − γ) = 2(1 − h) 1 − h (1 − FXe (0))g 1 − h #   √ (α − γ)(1 − FXe (γ)) −1 γ−α + 1−h . − Φ −Yg,h ς 1 − FXe (0) γ−α ς   √ 1−h  (24) Proof. See Appendix A. Lemma 4.2(ii) allows us to efficiently generate random samples from the severity distribution FX , thus allowing us to approximate the distribution of quantities of interest in Example 3.3 via Monte Carlo. Lemma 4.2(iii) shows that the Assumption (i) in Remark 3.4 is satisfied as long as the expected value of the frequency distribution, i.e. E[N ], is also tractable. Lemma 4.2(i) provides the distribution function that can be used to approximate the distribution function of L(d, W ). Concretely, by adopting the fast Fourier transform (FFT) approach with exponential tilting (see e.g. Embrechts and Frei (2009); Cruz et al. (2015)), we approximate the distribution function of L(d, W ), denoted by FL(d,W ) , by a finitely P (d)  (d) supported discrete distribution FbL(d,W ) (x) = j∈A pj ✶{a(d) ≤x} , where aj j∈A ⊂ R+ is a finite set j (d)  of atoms and pj j∈A are the corresponding probabilities. The details of the FFT approach with expo- nential tilting are shown in Algorithm 2. After obtaining (FbL(d,W ) )d∈D from Algorithm 2, the quantities h   +i and P λBM (b, t, L(d, W )) ∈ I in Remark 3.4 can E ✶I (λBM (b, t, L(d, W ))) λBM (b, t, L(d, W )) − α be approximated by finite sums: h + +i X (d) (d) (d)  BM λ (b, t, aj ) − α , pj ✶I λBM b, t, aj E ✶I (λBM (b, t, L(d, W ))) λBM (b, t, L(d, W )) − α ≈ j∈A   P λBM (b, t, L(d, W )) ∈ I ≈ X j∈A  . ✶I λBM b, t, a(d) j (d) pj One may increase the granularity parameter Kgr in Algorithm 2 to increase the precision of numerical approximation. Consequently, Assumptions (ii) and (iii) in Remark 3.4 are satisfied, and hence, Algorithm 1 is tractable and efficient in this setting. In particular, Algorithm 2 only needs to be executed once before executing Algorithm 1. 16 Algorithm 2: Fast Fourier Transform Approach with Exponential Tilting for Approximating FL(d,W ) (see Embrechts and Frei (2009)) Input: D, FX (·), ψN (·), γ(·), l, Kgr ∈ N, θ > 0 P (d) (d) (d) Output: (aj , pj )j∈A,d∈D , FbL(d,W ) (x) = j∈A pj ✶{a(d) ≤x} for each d ∈ D j 1 2 ε ← (2Kgr − 1)−1 l, A ← {0, 1, . . . , 2Kgr − 1}. for d ∈ D do (d) ← jǫ for each j ∈ A.   ← exp(−jθ) FX,d (jε + 21 ε) − FX,d (jε − 12 ε) for each j ∈ A, where  FX,d (y) := FX y + γ(d) ✶{y≥0} . P (d) (d) ϕj ← k∈A exp(iπ21−Kgr jk)fk for each j ∈ A via the FFT algorithm. aj 3 (d) fj 4 5 (d) 6 ψj 7 pj (d) (d) ← ψN (ϕj ) for each j ∈ A. P (d) ← exp(jθ)2−Kgr k∈A exp(−iπ21−Kgr jk)ψk for each j ∈ A via the inverse FFT algorithm. 8 P (d) (d) (d) return (aj , pj )j∈A,d∈D , FbL(d,W ) (x) = j∈A pj ✶{a(d) ≤x} for each d ∈ D. j 5 Numerical Experiments In Section 3 and Section 4, we formulated the optimal cybersecurity provisioning problem as a finite horizon stochastic optimal control problem, and developed a dynamic programming algorithm, i.e. Algorithm 1, to efficiently solve the problem under the assumption that the loss severity follows the truncated g-and-h distribution. Algorithm 1 not only computes the optimal cybersecurity provisioning policy for the insured, but also computes related quantities of interest, such as those in Example 3.3, that can guide the insurer when designing a suitable cyber risk insurance contract with a Bonus-Malus system. In this section, we demonstrate how Algorithm 1 aids the insurer when designing a cyber risk insurance contract and the benefits of the Bonus-Malus system by a numerical experiment.5 In particular, we investigate two aspects of the cyber risk insurance contract with Bonus-Malus. The first aspect is whether the presence of the cyber risk insurance contract dis-incentivizes the adoption of self-mitigation measures, an issue known as moral hazard. The second aspect is whether the Bonus-Malus system provides benefits to the insurer in terms of increased customer retention rates and discounted expected profits. 5.1 Experimental Settings We assume that all monetary quantities, including the severity of cyber risk events, the insurance premium, and the annual investment required by self-mitigation measures are adjusted to the scale of the insured (e.g. its average annual revenue) and are unit-free. We consider insurance policies that last for 20 years, that is, T = 20. The discount factor e−r is fixed at 0.95. In the cyber loss model, we let the frequency distribution be the Poisson distribution with rate 0.8. We set the severity distribution to be Tr-g-and-h(α = 0, ς = 1, g = 1.8, h = 0.15), where the g and h parameters are set to be similar to those 5 The code used in this work for the experiment is available on GitHub: CyberInsuranceBonusMalus 17 https://github.com/qikunxiang/ estimated in Dutta and Perry (2006) from real Operational Risk data (see Table 8 of Dutta and Perry (2006)). We would like to remark that the heaviness of the tail of the loss severity distribution (i.e. the parameter h in the truncated g-and-h distribution) determines the probability of extreme risk events and is crucial in the computation of capital estimate (Dutta and Perry, 2006). Therefore, it is important that we specify a realistic value of the parameter h. In Algorithm 2, we fix l = 10000, Kgr = 20, θ= 20 2Kgr = 3.0518 × 10−4 . For simplicity, we consider the situation where only a single self-mitigation measure is available, that is, D = 1. This self-mitigation measure requires an annual investment of 0.5, and has the effect of preventing 70% of the incoming cyber risk events and decreasing the severity of the remaining events −1 by the 70th percentile of the severity distribution, that is, β(1) = 0.5, γ(1) = FX (0.7), where X ∼ Tr-g-and-h(α = 0, ς = 1, g = 1.8, h = 0.15). We consider the following simple cyber risk insurance policy with Bonus-Malus system. Let B = {−2, −1, 0, 1}, and let the functions BM(bt−1 , Ct ) and BM0 (bt−1 , it−1 ) be specified in Table 5.1 below. Table 1: The BM(·, ·) and BM0 (·, ·) functions that represent the Bonus-Malus update rules BM(bt−1 , Ct ) bt−1 Ct =0 >0 −2 −2 1 −1 −2 1 0 −1 1 1 0 1 BM0 (bt−1 , it−1 ) bt−1 it−1 on off1 −2 (−2, off1 ) (−1, off1 ) −1 (−1, off1 ) (0, off1 ) 0 (0, off1 ) (0, off1 ) 1 (1, off1 ) (0, off1 ) The above settings mean that when the contract is activated, the insured is migrated to level 1 in the following policy year whenever a claim is made. When the insured does not make any claim in a policy year, their policy is migrated downwards by one level in the following policy years until it reaches level −2. When the contract is deactivated, if the insured’s policy is in level 1, it is migrated back to level 0 after one year. Otherwise, the policy is migrated upwards by one level each year until it reaches level 0. In the experiment, we let the base premium pBM base be an adjustable parameter that is varied between 0 and 7 with an increment of 0.005, and set the premium to be 60%, 80%, 100%, 150% of the base premium for Bonus-Malus levels −2, −1, 0, 1, respectively. That is, we let pBM (−2, t) = 0.6pBM base , BM BM pBM (−1, t) = 0.8pBM (0, t) = pBM (1, t) = 1.5pBM base , p base , p base for all t ∈ {1, . . . , T }. We fix the BM maximum compensation to be 1000, that is, lmax (b, t) = 1000 for all b ∈ B, t ∈ {1, . . . , T }. We set the deductible to be 0.5 for all but the last policy year, and set the deductible to be 5 for the last policy year, BM BM that is, ldtb (b, t) = 0.5 for all b ∈ B, t ∈ {1, . . . , T − 1} and ldtb (b, T ) = 5 for all b ∈ B. This is to prevent an issue caused by the finite horizon. Since after the last policy year there is no future benefit from the insurance policy and the insured is not incentivized to adopt the self-mitigation measure, a higher deductible is used as the incentive in the last policy year. In addition, we let δin (t) = 0.75(t − 16)+ , δout (t) = 3 + 5 19 (t − 1), and δre = 3. This setting has the effect of incentivizing the insured to activate 18 Without Bonus-Malus With Bonus-Malus 20 years years 20 10 uninsured insured 0 0 0 1 2 3 4 5 6 7 0 base premium 1 2 10 without mitigation with mitigation 0 3 4 5 6 7 base premium 20 years 20 years uninsured level -2 level -1 level 0 level 1 10 10 without mitigation with mitigation 0 0 1 2 3 4 5 6 7 0 1 base premium 2 3 4 5 6 7 base premium Figure 1: The retention of the cyber risk insurance policy and the expected years of adoption of the self-mitigation measure versus the base premium. the insurance contract early on, and dis-incentivizing withdrawal when close to the last policy year. As a baseline for comparison, we also consider another cyber risk insurance policy without the Bonus-Malus system, which can be modeled by letting B = {0}. We fix the premium to be the base premium pBM base , and leave everything else identical to the policy with the Bonus-Malus system. 5.2 Results and Discussion Figure 1 shows the expected number of years the insured’s policy spends in each of the Bonus-Malus levels or being de-activated (uninsured) and the expected number of years the insured adopts the selfmitigation measure. The two panels compare the cyber risk insurance policy with the Bonus-Malus system with the one without. With the policy that does not have the Bonus-Malus system, the decisions of the insured are completely deterministic, that is, they do not depend on the realization of losses. When pBM base ≤ 4.410, the optimal strategy of the insured is to purchase the cyber risk insurance every year and only adopt the self-mitigation measure in the last policy year (due to the higher deductible in the last policy year). When pBM base ≥ 4.415, the optimal strategy of the insured is to never purchase the cyber risk insurance and always adopt the self-mitigation measure. Therefore, without the Bonus-Malus system, the issue of moral hazard is present and the insured will treat the cyber risk insurance and the self-mitigation measure as substitute goods. On the other hand, when the Bonus-Malus system is introduced to the cyber risk insurance policy, the decisions of the insured depend on the realization of losses. When 4.495 ≤ pBM base ≤ 4.930, the optimal strategy of the insured is to always purchase the cyber risk insurance and adopt the self-mitigation measure. When 4.935 ≤ pBM base ≤ 5.050, the optimal strategy of the insured is to always adopt the self-mitigation measure but withdraw from the contract when the expected future cost exceeds the expected future benefit of the insurance policy. As a result, the retention rate, i.e. the expected proportion of years the insured activates the contract, drops when the base premium is increased. When pBM base ≥ 5.055, the optimal strategy of the insured is to never purchase the cyber risk insurance and always adopt the self-mitigation measure. Hence, compared with the policy without Bonus-Malus, the policy with Bonus-Malus incentivizes the insured to adopt the self-mitigation measure in addition to purchasing the cyber risk insurance policy. Figure 2 compares both the expected value of the discounted total loss prevented by the self-mitigation measure and the expected value of the discounted profit of the insurer (defined in Section 3.3) in the 19 prevented loss 10 0 0 1 2 3 4 5 6 7 base premium 0 insurer's profit prevented loss insurer's profit Without Bonus-Malus 20 -50 0 1 2 3 4 5 6 7 With Bonus-Malus 20 10 0 0 2 3 4 5 6 7 5 6 7 base premium -50 0 base premium 1 0 1 2 3 4 base premium Figure 2: The discounted total expected loss prevented by the self-mitigation measure and the discounted expected profit of the insurer versus the base premium. Left panel: the policy without the Bonus-Malus system. The dashed line indicates the highest base premium before the insured chooses not to purchase cyber risk insurance. Right panel: the policy with the Bonus-Malus system. The dashed line indicates the highest base premium before the retention drops below 100%. The dotted line indicates the highest base premium before the insured chooses not to purchase cyber risk insurance. two policies. The left panel of Figure 2 shows the case without Bonus-Malus. In that case, when pBM base ≤ 4.410, the insured will always purchase the cyber risk insurance policy but will only adopt the self-mitigation measure in the last policy year. Hence, the discounted total expected loss prevented stays at 0.505, while the discounted expected profit of the insurer increases as the base premium increases. When pBM base ≥ 4.415, the insured will not purchase the insurance policy but will always adopt the selfmitigation measure. As a result, the discounted total expected loss prevented will be 17.183 but the insurer will earn no profit. The most the insurer can gain before losing the insured is −10.510, when the base premium is set to 4.410. In contrast, in the case with the Bonus-Malus system, as shown in the right panel of Figure 2, the insurer can gain a discounted expected profit of at most −0.860 while always retaining the insured (i.e. the insured will never withdraw from the contract), when the base premium is set to 4.930. The insurer can gain a discounted expected profit of at most −0.006 before losing the insured, when the base premium is set to 5.050. With both of these base premiums, the insured will always adopt the self-mitigation measure, resulting in a discounted total expected loss prevention of 17.183. Overall, this experiment demonstrates two benefits of the Bonus-Malus system. First, the presence of the Bonus-Malus system incentivizes the insured to adopt the self-mitigation measure in addition to the cyber risk insurance policy. This results in a considerable increase in the prevention of cyber losses, which enhances the overall security of the cyberspace. Second, the Bonus-Malus system benefits the insurer, since it allows the insurer to gain more profit from the cyber risk insurance policy while remaining attractive to the insured. 6 Conclusion This paper motivated the joint consideration of risk reduction and risk transfer decisions in the face of cyber risk. We introduced a cyber risk insurance policy with a Bonus-Malus system to provide incentive 20 mechanisms to promote the adoption of cyber risk mitigation practices. We developed a model based on the stochastic optimal control framework to analyze how a rational insured allocates funds between risk mitigation measures and the cyber risk insurance policy. A dynamic programming-based algorithm was then developed to efficiently solve this decision problem. A numerical experiment demonstrated that this novel type of insurance policy can incentivize the adoption of risk mitigation measures and can allow the insurer to profit more from the policy while remaining attractive to the insured. Future research could investigate the effects of the risk profile, i.e. the characteristics of the loss distribution such as the heaviness of its tail, on the effectiveness of the Bonus-Malus system and how one can tailor Bonus-Malus-based insurance contracts for different risk profiles. Acknowledgments Ariel Neufeld gratefully acknowledges the financial support by his Nanyang Assistant Professorship Grant (NAP Grant) Machine Learning based Algorithms in Finance and Insurance. A Proofs Proof of Theorem 3.1. In this proof, we apply dynamic programming and perform backward induction in time to show the optimality of π ⋆ . First, one may check that d⋆t , ι⋆t are Ft−1 -measurable, jt⋆ is Ft measurable, and {ι⋆t = 0, jt⋆ = 1} = ∅ for t = 1, . . . , T . Thus, indeed π ⋆ ∈ Π. For all π = (ds , ιs , js )s=1:T ∈ Π and t ∈ {0, . . . , T }, let us define Ot (π) = (des , e ιs , e js )s=1:T ∈ Π as follows: O (π) O (π)  b0 t , i 0 t := (0, no), for each s =1, . . . , t, let: des = ds , e ιs = ιs , for each s =t + 1, . . . , T, let: Ot (π) Ot (π)  des = dbs bs−1 , is−1 , O (π)  O (π) t t e ιs = b ιs bs−1 , is−1 e js = js , , (25)  Ot (π) Ot (π) e js = b js bs−1 , is−1 , Ws . By the definition above, one may check that Ot (π) ∈ Π for all π ∈ Π and t = 0, . . . , T . In particular,  when t = 0, (25) implies that O0 (π) = π ⋆ for all π ∈ Π. In addition, notice that Ot+s Ot (π) = Ot (π) for all π ∈ Π and s ≥ 0. Next, we prove the following statement by induction: for all t = 0, . . . , T, Ot (π) V t bt Ot (π)  , it Ot (π) = Vt ≤ Vtπ P-a.s. for all π ∈ Π. (26) To begin, we have by definition that OT (π) = π and VT (bπT , iπT ) = VTπ = 0 P-a.s. for all π ∈ Π. Hence, (26) O (π) O (π)  = holds when t = T . Now, let us suppose that for some t ∈ {1, . . . , T }, it holds that Vt bt t , it t Ot (π) Vt ≤ Vtπ P-a.s. for all π ∈ Π. Let π = (ds , ιs , js )s=1:T ∈ Π be arbitrary and let Ot−1 (π) := (des , e ιs , e js )s=1:T ∈ Π. Note that (12) ensures that for every b ∈ B, i ∈ I, w ∈ W, gt (b, i, dbt (b, i), b ιt (b, i), b jt (b, i, w), w) + Vt ft (b, i, dbt (b, i), b ιt (b, i), b jt (b, i, w), w) n o = min gt (b, i, dbt (b, i), b ιt (b, i), j, w) + Vt ft (b, i, dbt (b, i), b ιt (b, i), j, w) . j∈{0,1} 21  (27) Combining (27) with the definition of dbt and b ιt in (11), one can show that for all b ∈ B, i ∈ I, d ∈ D, ι ∈ {0, 1}, and B(W)-measurable j : W → {0, 1}, it holds that Vt−1 (b, i)    −r b b b b =e E gt (b, i, dt (b, i), b ιt (b, i), jt (b, i, W ), W ) + Vt ft (b, i, dt (b, i), b ιt (b, i), jt (b, i, W ), W )    ≤e−r E gt (b, i, d, ι, j(W ), W ) + Vt ft (b, i, d, ι, j(W ), W ) . (28) By (28), (25), the independence between Ft−1 and σ(Wt ), and the induction hypothesis, it holds P-a.s. that, Ot−1 (π) Ot−1 (π)  Vt−1 bt−1 , it−1  =e−r E gt (b, i, dbt (b, i), b ιt (b, i), b jt (b, i, W ), W )  + Vt ft (b, i, dbt (b, i), b ιt (b, i), b jt (b, i, W ), W ) O t−1 =e−r E gt (bt−1 (π) O t−1 , it−1 (π)   O t−1 b=bt−1 (π) O t−1 , i=it−1 (π) Ot−1 (π) Ot−1 (π) Ot−1 (π) Ot−1 (π) , dbt (bt−1 , it−1 ), b ιt (bt−1 , it−1 ), Ot−1 (π) Ot−1 (π) Ot−1 (π) Ot−1 (π) b Ot−1 (π) Ot−1 (π) b , it−1 , dt (bt−1 , it−1 ), jt (bt−1 , it−1 , Wt ), Wt ) + Vt ft (bt−1   Ot−1 (π) Ot−1 (π) b Ot−1 (π) Ot−1 (π) b ιt (bt−1 , it−1 ), jt (bt−1 , it−1 , Wt ), Wt ) Ft−1 (29)     Ot−1 (π) Ot−1 (π) e Ot−1 (π) Ot−1 (π) e , it−1 , dt , e ιt , e jt , Wt Ft−1 =e−r E gt bt−1 , it−1 , dt , e ιt , e jt , Wt + Vt ft bt−1    Ot−1 (π) Ot−1 (π) e O (O (π)) Ot (Ot−1 (π))  =e−r E gt bt−1 , it−1 , dt , e ιt , e jt , Wt + Vt bt t t−1 , it Ft−1    Ot−1 (π) Ot−1 (π) e O (π) =e−r E gt bt−1 , it−1 , dt , e ιt , e jt , Wt + Vt t−1 Ft−1 O =Vt−1t−1 (π) . Now, let π = (ds , ιs , js )s=1:T ∈ Π be arbitrary and let Ot−1 (π) := (des , e ιs , e js )s=1:T ∈ Π. By (25), we have that  Ot−1 (π) Ot−1 (π)  Ot (π) Ot (π)  , it−1 = bt−1 , it−1 bπt−1 , iπt−1 = bt−1 P-a.s. (30) By (29), (27), (30), the independence between Ft−1 and σ(Wt ), and the induction hypothesis, we have 22 P-a.s. that (π) O Vt−1t−1   n o −r b b =e E min gt (b, i, dt (b, i), b ιt (b, i), j, W ) + Vt ft (b, i, dt (b, i), b ιt (b, i), j, W ) j∈{0,1} ≤e−r E   min j∈{0,1} n n gt (b, i, dt , ιt , j, W ) + Vt ft (b, i, dt , ιt , j, W )  o O t−1 b=bt−1 (π) O t−1 , i=it−1 O t−1 b=bt−1 (π) O t−1 , i=it−1 (π)  o Ot−1 (π) Ot−1 (π) , dt , ιt , j, Wt + Vt ft bt−1 , it−1 , dt , ιt , j, Wt Ft−1 j∈{0,1}     Ot−1 (π) Ot−1 (π) Ot−1 (π) Ot−1 (π) ≤e−r E gt bt−1 , it−1 , dt , ιt , jt , Wt + Vt ft bt−1 , it−1 , dt , ιt , jt , Wt Ft−1    O (π) O (π)  =e−r E gt bπt−1 , iπt−1 , dt , ιt , jt , Wt + Vt bt t , it t Ft−1    ≤e−r E gt bπt−1 , iπt−1 , dt , ιt , jt , Wt + Vtπ Ft−1 =e−r E min O t−1 gt bt−1 (π) O t−1 , it−1 (π) (π)  π =Vt−1 . (31) O (π) Ot−1 (π)  π = Vt−1t−1 ≤ Vt−1 P-a.s. for all , it−1  ⋆ ⋆ ⋆ π ∈ Π. By induction, (26) holds for t = 0. Hence, V0 bπ0 , iπ0 = V0π ≤ V0π for all π ∈ Π. Since ⋆ ⋆ ⋆ ⋆ ⋆ (b0π , iπ0 = (0, no), we have V0 bπ0 , iπ0 = V0π = inf π∈Π V0π = V0 . The proof is now complete. O t−1 Combining (29) and (31), we have shown that Vt−1 bt−1 (π) Proof of Theorem 3.2. To prove statement (i), it suffices to show that the two following statements hold: (i-a) For all b ∈ B, i ∈ I, d ∈ D, ι ∈ {0, 1},   n o E min gt (b, i, d, ι, j, W ) + Vt ft (b, i, d, ι, j, W ) j∈{0,1} =β(d) + ιpBM (b, t) + δin (t)✶{i=no,ι=1} + δout (t)✶{i=on,ι=0}   + δre ✶{i6=on,i6=no,ι=1} + E L(d, W ) + Ht (b, i, d, ι), where Ht (b, i, d, ι) is defined on Line 9 and Line 10 of Algorithm 1. (i-b) For all b ∈ B, i ∈ I, d ∈ D,   gt (b, i, d, 1, 1, w) + Vt ft (b, i, d, 1, 1, w) < gt (b, i, d, 1, 0, w) + Vt ft (b, i, d, 1, 0, w) m λ BM (b, t, L(d, w)) ∈ [ b≤b′ ≤b (32) ′ Lt (b, b ), where b, b are defined on Line 4 of Algorithm 1, and Lt (b, b′ ) is defined on Line 6 of Algorithm 1. If statements (i-a) and (i-b) hold, then one can verify that dbt (b, i), b ιt (b, i), b jt (b, i, w) defined on Line 11 and Line 13 coincide with the definitions (11) and (12), and thus statement (i) holds as a consequence of Theorem 3.1. 23 In statement (i-a), in the case where ι = 1, we have, by (5) and (6), that   n o E min gt (b, i, d, ι, j, W ) + Vt ft (b, i, d, ι, j, W ) h j∈{0,1} =E β(d) + ιpBM (b, t) + δin (t)✶{i=no,ι=1} + δout (t)✶{i=on,ι=0} + δre ✶{i6=on,i6=no,ι=1} + L(d, W )  n o  BM BM + E min Vt BM(b, jλ (b, t, L(d, W ))), on − ιjλ (b, t, L(d, W )) i j∈{0,1}   =β(d) + ιpBM (b, t) + δin (t)✶{i=no,ι=1} + δout (t)✶{i=on,ι=0} + δre ✶{i6=on,i6=no,ι=1} + E L(d, W ) h i    + E Vt BM(b, 0), on ∧ Vt BM(b, λBM (b, t, L(d, W ))), on − λBM (b, t, L(d, W )) . Moreover, by the definition of Ht (b, i, d, ι) on Line 9, the definitions of b, b on Line 4, and the definition of αt (b, b′ ) on Line 6, we have h i    E Vt BM(b, 0), on ∧ Vt BM(b, λBM (b, t, L(d, W ))), on − λBM (b, t, L(d, W ))  =Vt BM(b, 0), on    + BM BM −E λ (b, t, L(d, W )) − Vt BM(b, λ (b, t, L(d, W ))), on + Vt BM(b, 0), on =Vt (b, on)   X   +  − E ✶{BM(b,λBM (b,t,L(d,W )))=b′ } λBM (b, t, L(d, W )) − Vt (b′ , on) − Vt (b, on) b≤b′ ≤b =Ht (b, i, d, 1). Thus, statement (i-a) holds when ι = 1. In the case where ι = 0, statement (i-a) follows directly from (5) and (6). In statement (i-b), by (5) and (6), it holds that   gt (b, i, d, 1, 1, w) + Vt ft (b, i, d, 1, 1, w) < gt (b, i, d, 1, 0, w) + Vt ft (b, i, d, 1, 0, w) m   V BM(b, λBM (b, t, L(d, w)), on − λBM (b, t, L(d, w) < V BM(b, 0), on . Observe that, by the definitions of b, b on Line 4 and the definition of αt (b, b′ ) on Line 6, we have n  o c ∈ R+ : Vt BM(b, c), on − c < Vt BM(b, 0), on o [ n c ∈ R+ : BM(b, c) = b′ , c > Vt (b′ , on) − Vt (b, on) = = b≤b′ ≤b [ b≤b′ ≤b Lt (b, b′ ). Hence, statement (i-b) holds. Now, let us prove statement (ii). Let (b, i) ∈ B × I be fixed, let b, b be defined by Line 4, and let ιt (b, i) = 0, then Lt (b, b′ ) be defined by Line 6. It follows from (5) that, if b  ⋆ ⋆    ⋆ ⋆ P bπt , iπt = BM0 (b, i) bπt−1 = b, iπt−1 = i = 1 = Pt⋆ (b, i) → BM0 (b, i) , thus showing the correctness of Line 20. Now, suppose that b ιt (b, i) = 1. Then, by (5), we have that 24   ⋆ ⋆ ⋆ P iπt = on bπt−1 = b, iπt−1 = i = 1. Let us first examine the case where b′ 6= b. We then have n o ⋆ ⋆ ⋆ ⋆  bπt , iπt = (b′ , on), bπt−1 , iπt−1 = (b, i) n   o o  n ⋆ ⋆  = BM b, λBM (b, t, L(d⋆t , Wt )) = b′ ∩ jt⋆ = 1 ∩ bπt−1 , iπt−1 = (b, i) o n   o n ⋆ o n ⋆  = BM b, λBM (b, t, L(dbt (b, i), Wt )) = b′ ∩ b jt (b, i, Wt ) = 1 ∩ bπt−1 , iπt−1 = (b, i) n   o n o S = BM b, λBM (b, t, L(dbt (b, i), Wt )) = b′ ∩ λBM (b, t, L(dbt (b, i), Wt )) ∈ b≤b′′ ≤b Lt (b, b′′ ) n ⋆ o ⋆  ∩ bπt−1 , iπt−1 = (b, i) o n o n ⋆ ⋆  = λBM (b, t, L(dbt (b, i), Wt )) ∈ Lt (b, b′ ) ∩ bπt−1 , iπt−1 = (b, i) , where the first equality is by (5), the second equality is by (13), the third equality is by statement (ib), and the last equality is by Line 6 and the property that {Lt (b, b′′ ) : b ≤ b′′ ≤ b} are disjoint sets. ⋆  ⋆ Therefore, since bπt−1 , iπt−1 and Wt are independent, we have for any b′ 6= b that i h ⋆ i h i h ⋆ ⋆ ⋆ ⋆ ⋆ P bπt = b′ , iπt = on, bπt−1 = b, iπt−1 i = P λBM (b, t, L(dbt (b, i), Wt )) ∈ Lt (b, b′ ) P bπt−1 = b, iπt−1 = i . Hence, by Line 17, h ⋆ i h i ⋆ ⋆ ⋆ P bπt = b′ , iπt = on bπt−1 = b, iπt−1 = i =P λBM (b, t, L(dbt (b, i), Wt )) ∈ Lt (b, b′ )   =Pt⋆ (b, i) → (b′ , on) . The remaining case where b′ = b follows from i h ⋆ ⋆ i X h ⋆ ⋆ ⋆  ⋆  ⋆ ⋆ P bπt , iπt = (b, on) bπt−1 , iπt−1 = (b, i) =1 − P bπt , iπt = (b′ , on) bπt−1 , iπt−1 = (b, i) b<b′ ≤b =1 − X b<b′ ≤b   Pt⋆ (b, i) → (b′ , on)   =Pt⋆ (b, i) → (b, on) , thus verifying the correctness of Line 18 of Algorithm 1. This completes the proof of (15). Equation (16) ⋆ ⋆ follows from the definition that bπ0 , iπ0 = (0, no) and basic properties of a finite state Markov chain. The proof of statement (ii) is complete. Finally, statement (iii) also follows from the basic properties of a finite state Markov chain. The proof is complete. Proof of Lemma 4.2. Statement (i) follows by checking the following:    FXf(x)−FXf(0) e 1−FX f(0) e ≤ x|X e > 0] = P[0 < X ≤ x] = FX (x) =P[X e  P[X > 0] 0 if x > 0, if x ≤ 0. Statement (ii) can be verified directly by checking that P[XU ≤ x] = FX (x) for all x ∈ R. 25 Finally, statement (iii) can be derived from (22) as follows:   E (X − γ)+ Z ∞ (x − γ)FX (dx) = γ  Z ∞ 1 = xFXe (dx) − γ(1 − FXe (γ)) 1 − FXe (0) γ Z ∞ (α − γ)(1 − FXe (γ)) ς = Yg,h (z)Φ(dz) + −1 γ−α 1 − FXe (0) Yg,h ( ς ) 1 − FXe (0)  2  2 Z ∞ (α − γ)(1 − FXe (γ)) ς hz 1 z 1 √ exp − = (exp(gz) − 1) exp dz + −1 γ−α 1 − FXe (0) g Yg,h 2 2 1 − FXe (0) 2π ( ς ) "Z  #    ∞ (α − γ)(1 − FXe (γ)) (1 − h)z 2 1 (1 − h)z 2 ς √ dz + + gz − exp − exp − = −1 γ−α 1 − FXe (0) g 2π Yg,h 2 2 1 − FXe (0) ( ς ) ! "Z     2 ∞ g2 (1 − h) g ς 1 √ exp dz = exp − z− −1 γ−α 1 − FXe (0) g 2π Yg,h 2(1 − h) 2 1−h ( ς )  #  Z ∞ (α − γ)(1 − FXe (γ)) (1 − h)z 2 dz + − exp − −1 γ−α 2 1 − FXe (0) Yg,h ( ς ) "       √ ς g2 g −1 γ−α √ 1−h exp Φ − Yg,h = ς 2(1 − h) 1−h (1 − FXe (0))g 1 − h #   √ (α − γ)(1 − FXe (γ)) −1 γ−α − Φ −Yg,h + 1−h , ς 1 − FXe (0) where the last equality is obtained by noticing that both integrals are Gaussian integrals after a change of variable. The proof is now complete. References Tridib Bandyopadhyay, Vijay S. Mookerjee, and Ram C. Rao. Why IT managers don’t go for cyberinsurance products. Commun. ACM, 52(11):68–73, November 2009. 3 Rainer Böhme and Galina Schwartz. Modeling cyber-insurance: Towards a unifying framework. In 9th Annual Workshop on the Economics of Information Security, WEIS 2010, Harvard University, Cambridge, MA, USA, June 7-8, 2010, 2010. 3 Jonathan Chase, Dusit Niyato, Ping Wang, Sivadon Chaisiri, and Ryan K. L. Ko. A scalable approach to joint cyber insurance and security-as-a-service provisioning in cloud computing. IEEE Transactions on Dependable and Secure Computing, 16(4):565–579, 2019. 4, 5 Dan Craigen, Nadia Diakun-Thibault, and Randy Purse. Defining cybersecurity. Technology Innovation Management Review, 4(10):13–21, 2014. 2 Marcelo G. Cruz, Gareth W. Peters, and Pavel V. Shevchenko. Fundamental aspects of operational risk and insurance analytics: A handbook of operational risk. John Wiley & Sons, 2015. 15, 16 26 Wanchun Dou, Wenda Tang, Xiaotong Wu, Lianyong Qi, Xiaolong Xu, Xuyun Zhang, and Chunhua Hu. An insurance theory based optimal cyber-insurance contract against moral hazard. Information Sciences, 527:576–589, 2020. 4, 5 Kabir Dutta and Jason Perry. A tale of tails: an empirical analysis of loss distribution models for estimating operational risk capital. Technical report, Federal Reserve Bank of Boston, 2006. 15, 18 Paul Embrechts and Marco Frei. Panjer recursion versus FFT for compound distributions. Math. Methods Oper. Res., 69(3):497–508, 2009. 16, 17 Matthias A. Fahrenwaldt, Stefan Weber, and Kerstin Weske. Pricing of cyber insurance contracts in a network model. Astin Bull., 48(3):1175–1218, 2018. 4 Shaohan Feng, Zehui Xiong, Dusit Niyato, Ping Wang, and Amir Leshem. Evolving risk management against advanced persistent threats in fog computing. In 2018 IEEE 7th International Conference on Cloud Networking (CloudNet), pages 1–6, 2018. 4, 5 Emilio Granados Franco. The Global Risks Report 2020, World Economic Forum. https://www. weforum.org/reports/the-global-risks-report-2020/, January 2020. Accessed: 2021-02-04. 1 B. B. Gupta and Omkar P. Badve. Taxonomy of dos and ddos attacks and desirable defense mechanism in a cloud computing environment. Neural Comput. Appl., 28(12):3655–3682, December 2017. 2 Dinh Thai Hoang, Ping Wang, Dusit Niyato, and Ekram Hossain. Charging and discharging of plug-in electric vehicles (pevs) in vehicle-to-grid (v2g) systems: A cyber insurance-based model. IEEE Access, 5:732–754, 2017. 4, 5 Martin Husák, Jana Komárková, Elias Bou-Harb, and Pavel Čeleda. Survey of attack projection, prediction, and forecasting in cyber security. IEEE Communications Surveys & Tutorials, 21(1):640–660, 2019. 2 Mohammad Mahdi Khalili, Parinaz Naghizadeh, and Mingyan Liu. Designing cyber insurance policies: The role of pre-screening and security interdependence. IEEE Transactions on Information Forensics and Security, 13(9):2226–2239, 2018. 4, 5 Xiao Lu, Dusit Niyato, Hai Jiang, Ping Wang, and H. Vincent Poor. Cyber insurance for heterogeneous wireless networks. IEEE Communications Magazine, 56(6):21–27, 2018a. 4 Xiao Lu, Dusit Niyato, Nicolas Privault, Hai Jiang, and Ping Wang. Managing physical layer security in wireless cellular networks: A cyber insurance approach. IEEE Journal on Selected Areas in Communications, 36(7):1648–1661, 2018b. 3, 4, 5 Thomas Maillart and Didier Sornette. Heavy-tailed distribution of cyber-risks. The European Physical Journal B, 75(3):357–364, 2010. 14 Angelica Marotta, Fabio Martinelli, Stefano Nanni, Albina Orlando, and Artsiom Yautsiukhin. Cyberinsurance survey. Computer Science Review, 24:35–61, 2017. 3 27 Steve Morgan. Cybercrime to cost the world $10.5 trillion annually by 2025. //cybersecurityventures.com/cybercrime-damage-costs-10-trillion-by-2025/, https: Novem- ber 2020. Accessed: 2021-02-04. 1 Ranjan Pal and Leana Golubchik. Analyzing self-defense investments in internet security under cyberinsurance coverage. In 2010 IEEE 30th International Conference on Distributed Computing Systems, pages 339–347, 2010. 4, 5 Ranjan Pal, Leana Golubchik, Konstantinos Psounis, and Pan Hui. Will cyber-insurance improve network security? a market analysis. In IEEE INFOCOM 2014 - IEEE Conference on Computer Communications, pages 235–243, 2014. 4, 5 Ranjan Pal, Leana Golubchik, Konstantinos Psounis, and Pan Hui. Security pricing as enabler of cyberinsurance a first look at differentiated pricing markets. IEEE Transactions on Dependable and Secure Computing, 16(2):358–372, 2019. 4 Gareth W. Peters and Pavel V. Shevchenko. Advances in heavy tailed risk modeling. Wiley Handbook in Financial Engineering and Econometrics. John Wiley & Sons, Inc., Hoboken, NJ, 2015. A handbook of operational risk. 15 Gareth W. Peters and Scott A. Sisson. Bayesian inference, Monte Carlo sampling and operational risk. Journal of Operational Risk, 1(3):27–50, December 2006. 15 Gareth W. Peters, Aaron D. Byrnes, and Pavel V. Shevchenko. Impact of insurance for operational risk: is it worthwhile to insure or be insured for severe losses? Insurance Math. Econom., 48(2):287–303, 2011. 3 Gareth W. Peters, Wilson Ye Chen, and Richard H. Gerlach. Estimating quantile families of loss distributions for non-life insurance modelling via L-moments. Risks, 4(2), 2016. 15 Gareth W. Peters, Pavel V. Shevchenko, and Ruben D. Cohen. Statistical machine learning analysis of cyber risk data: event case studies. In Diane Maurice, Jack Freund, and David Fairman, editors, FinTech: Growth and Deregulation, chapter 3. Risk Books, 2018a. 3 Gareth W. Peters, Pavel V. Shevchenko, and Ruben D. Cohen. Understanding cyber-risk and cyberinsurance. In Diane Maurice, Jack Freund, and David Fairman, editors, FinTech: Growth and Deregulation, chapter 12. Risk Books, 2018b. 2, 3 Thomas Rid and Peter McBurney. Cyber-weapons. The RUSI Journal, 157(1):6–13, 2012. 2 Galina A. Schwartz and S. Shankar Sastry. Cyber-insurance framework for large scale interdependent networks. In Proceedings of the 3rd International Conference on High Confidence Networked Systems, HiCoNS ’14, page 145–154, New York, NY, USA, 2014. Association for Computing Machinery. 5 Nikhil Shetty, Galina Schwartz, Mark Felegyhazi, and Jean Walrand. Competitive cyber-insurance and internet security. In Tyler Moore, David Pym, and Christos Ioannidis, editors, Economics of Information Security and Privacy, pages 229–247. Springer US, Boston, MA, 2010. 3 28 Jinal P. Tailor and Ashish D. Patel. A comprehensive survey: ransomware attacks prevention, monitoring and damage control. International Journal of Scientific Research, 4:2321–2705, 06 2017. 2 John W Tukey. Exploratory data analysis, volume 2. Reading, MA: Addison-Wesley, 1977. 15 Spencer Wheatley, Thomas Maillart, and Didier Sornette. The extreme risk of personal data breaches and the erosion of privacy. The European Physical Journal B, 89(1):1–12, 2016. 14 Maochao Xu and Lei Hua. Cybersecurity insurance: Modeling and pricing. North American Actuarial Journal, 23(2):220–249, 2019. 4 Yihuan Xu, Boris Iglewicz, and Inna Chervoneva. Robust estimation of the parameters of g-and-h distributions, with applications to outlier detection. Computational Statistics & Data Analysis, 75:66 – 80, 2014. 15 Zichao Yang and John C.S. Lui. Security adoption and influence of cyber-insurance markets in heterogeneous networks. Performance Evaluation, 74:1 – 17, 2014. 5 Adam Young and Moti Yung. Cryptovirology: extortion-based security threats and countermeasures. In Proceedings 1996 IEEE Symposium on Security and Privacy, pages 129–140, 1996. 2 Rui Zhang and Quanyan Zhu. Optimal cyber-insurance contract design for dynamic risk management and mitigation. Preprint arXiv:1804.00998, 2018. 5 Rui Zhang, Quanyan Zhu, and Yezekael Hayel. A bi-level game approach to attack-aware cyber insurance of computer networks. IEEE Journal on Selected Areas in Communications, 35(3):779–794, 2017. 5 29