Skip to main content
Log in

A general HJM framework for multiple yield curve modelling

  • Published:
Finance and Stochastics Aims and scope Submit manuscript

Abstract

We propose a general framework for modelling multiple yield curves which have emerged after the last financial crisis. In a general semimartingale setting, we provide an HJM approach to model the term structure of multiplicative spreads between FRA rates and simply compounded OIS risk-free forward rates. We derive an HJM drift and consistency condition ensuring absence of arbitrage and, in addition, we show how to construct models such that multiplicative spreads are greater than one and ordered with respect to the tenor’s length. When the driving semimartingale is an affine process, we obtain a flexible and tractable Markovian structure. Finally, we show that the proposed framework allows unifying and extending several recent approaches to multiple yield curve modelling.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. Artificial risky bonds have been introduced in a number of recent papers; see e.g. [9, 27, 32]. We want to emphasize that artificial risky bond prices are only introduced here as an explanatory tool and will not be considered in the following sections of the paper.

  2. Due to the deterministic short rate, it is not necessary to distinguish the expectations with respect to different measures, but we explicitly indicate them for consistency of the exposition with the general setting of the following sections.

  3. Note that for every tenor \(\delta\in\{\delta _{1},\ldots,\delta_{m}\}\), the process \(u_{i}^{\top}Y\) plays the role of the process \(Z\) appearing in (2.3).

  4. The definition of local exponent in [39, Definition A.6] is slightly different since it is defined in terms of the exponential compensator of \((\int_{0}^{t} \mathsf{i}\beta_{s} \,dX_{s})_{t\geq0}\), which is due to the fact that complex-valued processes are considered in that paper. Note also that since Itô semimartingales are quasi-left-continuous, the derivatives of the modified Laplace cumulant process and of the ordinary Laplace cumulant process coincide (see e.g. [40, Remarks on p. 408]).

  5. In particular, in [9] (according to the notation used therein), it holds that \(Z^{\delta}_{t}=\int_{t}^{t+\delta }g_{t}(s)\,ds\), for any tenor \(\delta>0\), with \(g_{t}(T)\) representing the spread between risky and risk-free instantaneous \(T\)-forward rates.

  6. Note that our multiplicative spread \(S^{\delta}(t,T)\) corresponds to \(K_{L}(t,T,T+\delta)\), according to the notation adopted in [43].

  7. Let us remark that in the presence of funding costs, absence of arbitrage is implied by the existence of an equivalent measure under which the risky assets \(S\) present in the market are (local) martingales when discounted with their corresponding funding rate \(r^{f}\). This can be embedded in the classical framework where ℚ is a risk neutral measure with the risk-free (OIS) bank account as numéraire, by treating \(e^{\int_{0}^{\cdot} (-r_{s}^{f}+r_{s}) \,ds}S\) as traded asset.

  8. Indeed, as explained in [56, Sect. 4.2.1], bond prices are expressed in units of the respective currencies, while discount factors are simply the corresponding real numbers.

References

  1. Ametrano, F.M., Bianchetti, M.: Everything you always wanted to know about multiple interest rate curve bootstrapping but were afraid to ask (2013). Preprint available at http://ssrn.com/abstract=2219548

  2. Bianchetti, M.: Two curves, one price. Risk Magazine, 74–80 (2010)

  3. Bielecki, T., Rutkowski, M.: Valuation and hedging of contracts with funding costs and collateralization. SIAM J. Financ. Math. 6, 594–655 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  4. Brace, A., Gątarek, D., Musiela, M.: The market model of interest rate dynamics. Math. Finance 7, 127–155 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  5. Brigo, D., Mercurio, F.: A deterministic-shift extension of analytically-tractable and time-homogeneous short-rate models. Finance Stoch. 5, 369–387 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  6. Bru, M.F.: Wishart processes. J. Theor. Probab. 4, 725–751 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  7. Carmona, R.A.H.: A unified approach to dynamic models for fixed income, credit and equity markets. In: Carmona, R.A., et al. (eds.) Paris–Princeton Lectures on Mathematical Finance 2004. Lecture Notes in Math., vol. 1919, pp. 1–50. Springer, Berlin (2007)

    Chapter  Google Scholar 

  8. Crépey, S., Grbac, Z., Nguyen, H., Skovmand, D.: A Lévy HJM multiple-curve model with application to CVA computation. Quant. Finance 15, 401–419 (2015)

    Article  MathSciNet  Google Scholar 

  9. Crépey, S., Grbac, Z., Nguyen, H.N.: A multiple-curve HJM model of interbank risk. Math. Financ. Econ. 6, 155–190 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  10. Cuchiero, C., Filipović, D., Mayerhofer, E., Teichmann, J.: Affine processes on positive semidefinite matrices. Ann. Appl. Probab. 21, 397–463 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  11. Cuchiero, C., Klein, I., Teichmann, J.: A new perspective on the fundamental theorem of asset pricing for large financial markets. Theory of Probability and its Applications. To appear. Preprint available at http://arxiv.org/abs/1412.7562

  12. Cuchiero, C., Teichmann, J.: Path properties and regularity of affine processes on general state spaces. In: Donati-Martin, C., et al. (eds.) Séminaire de Probabilités XLV. Lecture Notes in Mathematics, vol. 2078, pp. 201–244. Springer, Berlin (2013)

    Chapter  Google Scholar 

  13. Delbaen, F., Schachermayer, W.: A general version of the fundamental theorem of asset pricing. Math. Ann. 300, 463–520 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  14. Dörsek, P., Teichmann, J.: Efficient simulation and calibration of general HJM models by splitting schemes. SIAM J. Financ. Math. 4, 575–598 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  15. Douady, R., Jeanblanc, M.: A rating-based model for credit derivatives. Eur. Invest. Rev. 1, 17–29 (2002)

    Google Scholar 

  16. Duffie, D., Filipović, D., Schachermayer, W.: Affine processes and applications in finance. Ann. Appl. Probab. 13, 984–1053 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  17. Eberlein, E., Kluge, W.: Calibration of Lévy term structure models. In: Fu, M., et al. (eds.) Advances in Mathematical Finance: In Honor of D.B. Madan, pp. 147–172. Birkhäuser, Basel (2007)

    Chapter  Google Scholar 

  18. Eberlein, E., Koval, N.: A cross-currency Lévy market model. Quant. Finance 6, 465–480 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  19. Filipović, D.: Consistency Problems for Heath–Jarrow–Morton Interest Rate Models. Lecture Notes in Mathematics, vol. 1760. Springer, Berlin (2001)

    Book  MATH  Google Scholar 

  20. Filipović, D., Overbeck, L., Schmidt, T.: Dynamic CDO term structure modeling. Math. Finance 21, 53–71 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  21. Filipović, D., Tappe, S., Teichmann, J.: Jump-diffusions in Hilbert spaces: existence, stability and numerics. Stochastics 82, 475–520 (2010)

    MathSciNet  MATH  Google Scholar 

  22. Filipović, D., Tappe, S., Teichmann, J.: Term structure models driven by Wiener processes and Poisson measures: existence and positivity. SIAM J. Financ. Math. 1, 523–554 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  23. Filipović, D., Trolle, A.B.: The term structure of interbank risk. J. Financ. Econ. 109, 707–733 (2013)

    Article  Google Scholar 

  24. Fries, C.P.: Curves and term structure models: definition, calibration and application of rate curves and term structure models (2010). Preprint available at http://ssrn.com/abstract=2194907

  25. Fujii, M., Shimada, A., Takahashi, A.: A market model of interest rates with dynamic basis spreads in the presence of collateral and multiple currencies. Wilmott 54, 61–73 (2011)

    Article  Google Scholar 

  26. Glau, K., Grbac, Z., Kriens, D.: Martingale property of exponential semimartingales: a note on explicit conditions and applications to asset price and Libor models (2015). Preprint, TU Munich, available at http://arxiv.org/pdf/1506.08127v1.pdf

  27. Grasselli, M., Miglietta, G.: A flexible spot multiple-curve model (2014). Preprint available at http://ssrn.com/abstract=2424242

  28. Grbac, Z., Papapantoleon, A.: A tractable LIBOR model with default risk. Math. Financ. Econ. 7, 203–227 (2013)

    Article  MATH  Google Scholar 

  29. Grbac, Z., Papapantoleon, A., Schoenmakers, J., Skovmand, D.: Affine LIBOR models with multiple curves: theory, examples and calibration. SIAM J. Financ. Math. 6, 984–1025 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  30. Heath, D., Jarrow, R., Morton, A.: Bond pricing and the term structure of interest rates: a new methodology for contingent claims valuation. Econometrica 60, 77–105 (1992)

    Article  MATH  Google Scholar 

  31. Henrard, M.: The irony in the derivatives discounting. Wilmott 30, 92–98 (2007)

    Google Scholar 

  32. Henrard, M.: The irony in the derivatives discounting part II: the crisis. Wilmott 2, 301–316 (2010)

    Article  Google Scholar 

  33. Henrard, M.: Interest Rate Modelling in the Multi-Curve Framework. Palgrave Macmillan, Basingstoke (2014)

    Book  Google Scholar 

  34. Jacod, J.: Multivariate point processes: predictable projection, Radon–Nikodým derivatives, representation of martingales. Z. Wahrscheinlichkeitstheor. Verw. Geb. 31, 235–253 (1974/1975)

  35. Jacod, J., Protter, P.: Discretization of Processes, Stochastic Modelling and Applied Probability, vol. 67. Springer, Berlin (2012)

    MATH  Google Scholar 

  36. Jacod, J., Shiryaev, A.: Limit Theorems for Stochastic Processes, 2nd edn. Grundlehren der mathematischen Wissenschaften, vol. 288. Springer, Berlin/Heidelberg/New York (2003)

    Book  MATH  Google Scholar 

  37. Jarrow, R., Turnbull, S.: Pricing derivatives on financial securities subject to credit risk. J. Finance 50, 53–85 (1995)

    Article  Google Scholar 

  38. Jeanblanc, M., Yor, M., Chesney, M.: Mathematical Methods for Financial Markets. Springer, London (2009)

    Book  MATH  Google Scholar 

  39. Kallsen, J., Krühner, P.: On a Heath–Jarrow–Morton approach for stock options. Finance Stoch. 19, 583–615 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  40. Kallsen, J., Shiryaev, A.: The cumulant process and Esscher’s change of measure. Finance Stoch. 6, 397–428 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  41. Keller-Ressel, M., Mayerhofer, E.: Exponential moments of affine processes. Ann. Appl. Probab. 25, 714–752 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  42. Kenyon, C.: Post-shock short-rate pricing. Risk Magazine, 83–87 (2010)

  43. Kijima, M., Tanaka, K., Wong, T.: A multi-quality model of interest rates. Quant. Finance 9, 133–145 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  44. Klein, I., Schmidt, T., Teichmann, J.: When roll-overs do not qualify as numéraire: bond markets beyond short rate paradigms (2013). Preprint available at http://arxiv.org/abs/1310.0032

  45. Kluge, W.: Time-inhomogeneous Lévy processes in interest rate and credit risk models. Ph.D. thesis, University of Freiburg (2005). available at https://www.freidok.uni-freiburg.de/data/2090/

  46. Koval, N.: Time-inhomogeneous Lévy processes in cross-currency market models. Ph.D. thesis, University of Freiburg (2005). available at https://www.freidok.uni-freiburg.de/data/2041/

  47. Krein, M.G., Nudelman, A.A.: The Markov Moment Problem and Extremal Problems: Ideas and Problems of P.L. Čebyšev and A.A. Markov and Their Further Development. Translations of Mathematical Monographs, vol. 50. Am. Math. Soc., Providence (1977)

    Google Scholar 

  48. Larsson, M., Ruf, J.: Convergence of local supermartingales and Novikov–Kazamaki type conditions for processes with jumps (2014). Preprint available at http://arxiv.org/abs/1411.6229

  49. Levin, K., Zhang, J.X.: Bloomberg volatility cube. Tech. rep., Bloomberg L.P. (2014)

  50. Mercurio, F.: Modern Libor market models: using different curves for projecting rates and discounting. Int. J. Theor. Appl. Finance 13, 113–137 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  51. Mercurio, F.: A Libor market model with a stochastic basis. Risk Magazine, 96–101 (2013)

  52. Mercurio, F., Xie, Z.: The basis goes stochastic. Risk Magazine, 78–83 (2012)

  53. Moreni, N., Pallavicini, A.: Parsimonious HJM modelling for multiple yield-curve dynamics. Quant. Finance 14, 199–210 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  54. Morini, M.: Solving the puzzle in the interest rate market. In: Bianchetti, M., Morini, M. (eds.) Interest Rate Modelling After the Financial Crisis, pp. 61–106. Risk Books, London (2013)

    Google Scholar 

  55. Morino, L., Runggaldier, W.J.: On multicurve models for the term structure. In: Dieci, R., et al. (eds.) Nonlinear Economic Dynamics and Financial Modelling: Essays in Honour of Carl Chiarella, pp. 275–290. Springer, Berlin (2014)

    Google Scholar 

  56. Musiela, M., Rutkowski, M.: Martingale Methods in Financial Modelling, 2nd edn. Springer, Berlin (2005)

    MATH  Google Scholar 

  57. Pallavicini, A., Tarenghi, M.: Interest rate modelling with multiple yield curves (2010). Preprint available at http://ssrn.com/abstract=1629688

  58. Protter, P.E.: Stochastic Integration and Differential Equations. Stochastic Modelling and Applied Probability, vol. 21, 2nd edn. Springer, Berlin/Heidelberg/New York (2005). Version 2.1, corrected third printing

    Book  Google Scholar 

Download references

Acknowledgements

The research of the second author was partly supported by a Marie Curie Intra European Fellowship within the 7th European Community Framework Programme under grant agreement PIEF-GA-2012-332345.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christa Cuchiero.

Appendices

Appendix A: Pricing under collateral and FRA rates

Let us here briefly review pricing under perfect collateralization for general derivatives which we then apply to the pricing of FRAs. For a more detailed discussion on general valuation with collateralization and funding costs, we refer to the growing literature on this topic, e.g. [3] and the references therein. We here follow closely [23, Sect. 2.2].

Throughout, let \((\varOmega, \mathcal{F}, (\mathcal{F}_{t})_{t\geq0}, \mathbb {P})\) be a filtered probability space, where ℙ stands for the statistical/historical probability measure. We consider OIS zero coupon bonds as basic traded instruments, which play the role of risk-free zero coupon bonds in the classical setting. In order to guarantee no arbitrage, we assume that

  1. (i)

    There exists an OIS (risk-free) bank account denoted by \((B_{t})_{t\geq0}\) and such that \(B_{t}=\exp(\int_{0}^{t} r_{s} \,ds)\), where \(r\) denotes the OIS short rate;

  2. (ii)

    There exists an equivalent probability measure ℚ such that the OIS bonds for all maturities are ℚ-martingales when denominated in units of the OIS bank account.Footnote 7

Let now \(X\) be an \(\mathcal{F}_{T}\)-measurable payoff of some derivative security. We assume here a perfect collateral agreement where 100 % of the derivative’s present value \(V_{t}\) is posted in the collateral at any time \(t < T\). The receiver of the collateral can invest it at the risk-free rate \(r\), corresponding to the OIS short rate, and has to pay an agreed collateral rate \(r^{c}_{t}\) to the poster of the collateral. Applying risk neutral pricing, we obtain for the present value of the collateralized transaction the expression

$$V_{t}=\mathbb{E}^{\mathbb{Q}}\bigg[ e^{-\int_{t}^{T} r_{s} \,ds}X+\int_{t}^{T} e^{-\int_{t}^{s} r_{u} \,du} (r_{s} -r^{c}_{s})V_{s} \,ds \bigg| \mathcal{F}_{t}\bigg]. $$

As shown in [23, Appendix A], this formula is equivalent to

$$V_{t}=\mathbb{E}^{\mathbb{Q}}\Big[ e^{-\int_{t}^{T} r^{c}_{s} \,ds}X \Big| \mathcal {F}_{t}\Big]. $$

Assuming that the collateral rate \(r^{c}\) corresponds to the OIS short rate \(r\), which is usually the case, we obtain the classical risk neutral valuation formula.

Since market quotes of FRAs correspond to perfectly collateralized contracts, where the collateral rate \(r^{c}\) is assumed to be the OIS short rate \(r\), the above pricing approach is applied for the definition of FRA rates. As in classical interest rate theory, the FRA rate, denoted by \(L_{t}(T,T+\delta)\), is the rate \(K\) fixed at time \(t\) such that the value of the FRA contract, whose payoff at time \(T+\delta\) is given by \(\delta(L_{T}(T,T+\delta)- K)\), is 0. Therefore, it holds that, for all \(t\in[0,T]\) and \(T\geq0\),

$$\mathbb{E}^{\mathbb{Q}}\big[e^{-\int_{t}^{T+\delta} r_{s} \,ds}\big(L_{T}(T,T+\delta)-K\big) \big| \mathcal{F}_{t}\big] \stackrel{!}{=}0. $$

Hence by Bayes’ formula,

$$\begin{aligned} L_{t}(T,T+\delta)=\mathbb{E}^{\mathbb{Q}^{T+\delta}}[L_{T}(T,T+\delta) | \mathcal{F}_{t}], \end{aligned}$$

where \(\mathbb{Q}^{T+\delta}\) denotes the \((T+\delta)\)-forward measure associated with the numéraire \(B(\cdot,T+\delta)\) and density process \(\frac{d\mathbb{Q}^{T+\delta}}{d\mathbb {Q}}|_{\mathcal {F}_{t}}=\frac{B(t,T+\delta)}{B_{t} B(0,T+\delta)}\). In particular, this provides a rigorous justification for the market practice of taking the expression (2.1) as the definition of fair FRA rates.

Appendix B: Foreign exchange analogy

For simplicity of presentation, let us consider a fixed tenor \(\delta\) and define artificial “risky” bond prices \(B^{\delta}(t,T)\) at time \(t\) and maturity \(T\) for the tenor \(\delta\) by the relation, for all \(t\leq T\) and \(T\geq0\),

$$L_{t}(T,T+\delta)=:\frac{1}{\delta} \left(\frac{B^{\delta }(t,T)}{B^{\delta }(t,T+\delta)}-1\right). $$

While the family \(\{(B(t,T))_{t\in[0,T]},T\geq0\}\) represents prices of domestic risk-free bonds (in units of the domestic currency), \(\{ (B^{\delta}(t,T))_{t\in[0,T]},T\geq0\}\) can be thought of as representing prices of zero-coupon bonds of a foreign “risky” economy, expressed in units of the foreign currency.

According to this foreign exchange analogy, one is naturally led to look at the ratio

$$ R^{\delta}(t,T):=\frac{B(t,T)}{B^{\delta}(t,T)}, $$

for \(t\leq T\) and \(T\geq0\), where \(B^{\delta}(t,T)\) (resp., \(B(t,T)\)) has here to be seen as the discount factor for the foreign (resp., domestic) economy.Footnote 8 Note also that \(R^{\delta}(T,T)=1\), for all \(T\geq0\). Following the presentation in [56, Sect. 4.2.1], the quantity \(R^{\delta}(t,T)\) corresponds to the forward exchange premium between the domestic and the foreign currency over the time interval \([t,T]\). Indeed, in standard foreign exchange markets, there is the following no arbitrage relation between the spot exchange rate \(Q_{t}\) (domestic price of one unit of the foreign currency) and the forward exchange rate \(F(t,T)\) (forward price in domestic currency of one unit of the foreign currency paid at time \(T\)): we have

$$\frac{Q_{t}}{F(t,T)}=\frac{B(t,T)}{B^{\delta}(t,T)}=R^{\delta}(t,T). $$

The multiplicative spread \(S^{\delta}(t,T)\) introduced in (2.2) corresponds now to

$$S^{\delta}(t,T) = \frac{1+\delta L_{t}(T,T+\delta)}{1+\delta L^{D}_{t}(T,T+\delta)} =\frac{B^{\delta}(t,T)}{B(t,T)}\frac{B(t,T+\delta)}{B^{\delta }(t,T+\delta)} =\frac{R^{\delta}(t,T+\delta)}{R^{\delta}(t,T)}, $$

for all \(t\leq T\) and \(T\geq0\), while the spot multiplicative spread is simply given by

$$S^{\delta}(T,T)=R^{\delta}(T,T+\delta), $$

for all \(T\geq0\). In particular, note that \(R^{\delta}(T,T+\delta)\) corresponds exactly to the quantity \(Q^{\delta}_{T}\) considered in Sect. 2.1 which thus has the interpretation of a foreign exchange premium over \([T,T+\delta]\).

Since Libor rates reflect the overall credit risk of the Libor panel, the exchange rate premium \(R^{\delta}(t,T+\delta)\) can be seen as a market valuation (at time \(t\)) of the riskiness of the foreign economy, i.e. of the credit and liquidity quality of the current Libor panel over the period \([t,T+\delta]\). Moreover, according to the same interpretation, the quantity

$$S^{\delta}(t,T)=R^{\delta}(t,T+\delta)/R^{\delta}(t,T)=\mathbb {E}_{\mathbb{Q}^{T}}[R(T,T+\delta)|\mathcal{F}_{t}] $$

is thus an expectation of the riskiness of the future Libor panel over the future time period \([T,T+\delta]\), as seen from the market at time \(t\) (calculated as the relative riskiness of the current Libor panel over the period \([t,T+\delta]\) relative to the one over \([t,T]\)). For instance, a large value of \(S^{\delta}(t,T)\) would mean that the market anticipates a worsening of the credit quality of the Libor panel on \([T,T+\delta]\) compared to the credit quality on \([t,T]\) as seen at time \(t\).

In this sense, the multiplicative spread \(S^{\delta}(t,T)\) is a rather natural quantity to model in a multiple curve setting, because it represents the market’s expectation at time \(t\) (being computed from financial instruments traded at date \(t\)) of the credit and liquidity quality of the Libor panel over \([T,T+\delta]\).

Appendix C: Local independence and semimartingale decomposition

In this section, we let \((X,Y)\) be a general Itô semimartingale taking values in \(\mathbb{R}^{d+n}\) and denote by \(\varPsi^{X,Y}\) its local exponent and by \(\varPsi^{X}\) and \(\varPsi^{Y}\) the local exponents of \(X\) and \(Y\), respectively, and let \(\mathcal{U}^{X,Y}\) be defined as in Definition 3.2. In view of [39, Lemma A.11], the following definition is equivalent to the notion of local independence as given in [39, Definition A.10].

Definition C.1

We say that \(X\) and \(Y\) are locally independent if outside a \(d\mathbb{Q}\otimes dt\)-null set, it holds that

$$\varPsi_{t}^{X,Y}(u_{t},v_{t})(\omega)=\varPsi_{t}^{X}(u_{t})(\omega)+\varPsi _{t}^{Y}(v_{t})(\omega), \quad\text{for all }(u,v)\in\mathcal{U}^{(X,Y)}. $$

Following [39, Appendix A.3], let us recall the notion of semimartingale decomposition of \(Y\) relative to \(X\). We denote by \(c^{Y,X}\) and \(c^{X}\) the second local characteristic of \((Y,X)\) and \(X\), respectively, and by \(K^{Y,X}\) and \(K^{X}\) the third local characteristic of \((Y,X)\) and \(X\), respectively. Denote also by \(\mu^{Y,X}\) the jump measure of \((Y,X)\). Supposing that \(1\in\mathcal{U}^{Y}\) (i.e. \(Y\) is exponentially special, see Proposition 3.3), let

$$\begin{aligned} Y^{\parallel,i} := \log\mathcal{E}\bigg(&\int_{0}^{\cdot}\bigl(c_{t}^{Y^{i},X}(c_{t}^{X})^{-1}\bigr)\,dX^{c}_{t} \\ &{}+\int_{0}^{\cdot}\int (e^{y^{i}}-1)\textbf{1}_{\{ x\neq0\}}\bigl(\mu^{Y,X}(dy,dx,dt)-K^{Y,X}_{t}(dy,dx)\,dt\bigr)\bigg), \end{aligned}$$
(C.1)

for \(i=1,\ldots,n\), where \((c^{X})^{-1}\) denotes the pseudoinverse of the matrix \(c^{X}\) and \(X^{c}\) is the continuous local martingale part of \(X\) (see [36, Proposition I.4.27]). We call \(Y^{\parallel }:=(Y^{\parallel,1},\ldots,Y^{\parallel,n})^{\top}\) the dependent part of \(Y\) relative to \(X\) and \(Y^{\perp}:=Y-Y^{\parallel}\) the independent part of \(Y\) relative to \(X\). The following lemma corresponds to [39, Lemma A.22 and Lemma A.23].

Lemma C.2

Let \((X,Y)\) be an \(\mathbb{R}^{d\times n}\)-valued Itô semimartingale and suppose that \(1\in\mathcal{U}^{Y}\). Then the following hold:

  1. (i)

    \(Y\mapsto Y^{\parallel}\) is a projection, in the sense that \((Y^{\parallel})^{\parallel}=Y^{\parallel}\).

  2. (ii)

    If \(Z\) is an Itô semimartingale locally independent of \(X\), then \((Z+Y)^{\parallel}=Y^{\parallel}\).

  3. (iii)

    \(\exp(Y^{\parallel,i})\) is a local martingale, for all \(i=1,\ldots,n\).

  4. (iv)

    \(Y^{\perp}\) and \((Y^{\parallel},X)\) are locally independent semimartingales.

Appendix D: Proofs of the results of Sect. 4

Proof of Proposition 4.4

Let us fix any \(i\in\{0,1,\ldots,m\}\). By the same argument as in [19, Corollary 5.12], it can be shown that \(\kappa_{4}(H^{\lambda}_{m+1}) \subseteq H^{\lambda,0}_{1} \) and \(\kappa_{j}^{i}(H^{\lambda}_{m+1}) \subseteq H^{\lambda,0}_{1} \) for \(j=1,2\). In the sequel, \(C\) always denotes a positive constant which can vary from line to line. The following estimates can be derived similarly as in the proof of [22, Proposition 3.2] (to which we refer the reader for more details), by using Assumption 4.3, the Hölder inequality and [22, Theorem 2.1].

For \(\kappa_{3}^{i}(h)\) and \(\kappa_{5}^{i}(h)\), we have for all \(h \in H_{m+1}^{\lambda}\), \(\xi\in\mathbb{R}^{d}\) and \(s \in\mathbb{R}_{+}\) that

$$ |(\zeta^{i0}(h)(s))^{\top}\xi|\leq C \| \zeta^{i0}(h)\|_{\lambda ,d}\| \xi\|_{d} \quad\text{and}\quad |(\zeta^{0}(h)(s))^{\top}\xi| \leq C\| \zeta^{0}(h)\|_{\lambda,d}\| \xi\|_{d}, $$

and for all \(\xi\in\mathbb{R}^{d}\), \(\hat{\xi}\in\mathbb{R}^{n}\) and \(s \in\mathbb {R}_{+}\) that

$$\begin{aligned} \bigl|e^{u_{i}^{\top} \hat{\xi}+ (Z^{i0}(h)(s))^{\top}\xi}-1\bigr|&\leq Ce^{\|u_{i}\|_{n}\|\hat{\xi}\|_{n}+(C_{0}+C_{i}) \|\xi\|_{d}} \big(\|u_{i}\|_{n}\| \hat{\xi}\| _{n}+\| \zeta^{i0}(h)\|_{\lambda,d}\| \xi\|_{d}\big),\\ \bigl|e^{-(Z^{0} (h)(s))^{\top}\xi}-1\bigr|&\leq Ce^{C_{0} \|\xi\|_{d}}\| \zeta^{0}(h)\|_{\lambda,d}\| \xi\|_{d}. \end{aligned}$$

Together with (4.9) and (4.10), this shows that we have \(\lim_{s \to\infty}\kappa_{3}^{i}(h)(s)=0\) and \(\lim_{s \to\infty}\kappa_{5}^{i}(h)(s)=0\). Moreover, we have

$$\begin{aligned} &\int_{\mathbb{R_{+}}} \bigg(\int\Big(\big(\zeta^{i0}(h)(s)\big)^{\top }\xi\Big)^{2}\bigl(e^{u_{i}^{\top} \hat{\xi}+ (Z^{i0}(h)(s))^{\top }\xi}\bigr)K^{\hat{Y},X}(d\hat{\xi}, d\xi)\bigg)^{2} e^{\lambda s} \,ds \\ & \quad\leq C (M_{0}+M_{i})^{4}K_{i}^{2},\\ &\int_{\mathbb{R_{+}}} \bigg( \int\Big(\big(\zeta^{0}(h)(s)\big)^{\top}\xi \Big)^{2} e^{-(Z^{0} (h)(s))^{\top}\xi}F(d\xi)\bigg)^{2} e^{\lambda s} \,ds \leq CM_{0}^{4}K_{0}^{2},\\ &\int_{\mathbb{R_{+}}} \bigg(\int\frac{d}{ds}\big(\zeta ^{i0}(h)(s)\big)^{\top}\xi\bigl(e^{u_{i}^{\top} \hat{\xi}+ (Z^{i0}(h)(s))^{\top }\xi}-1\bigr)K^{\hat{Y},X}(d\hat{\xi}, d\xi)\bigg)^{2}e^{\lambda s} \,ds \\ & \quad\leq C (M_{0}+M_{i})^{2}K_{i},\\ &\int_{\mathbb{R_{+}}} \left(\int\frac{d}{ds}\bigl(\zeta ^{0}(h)(s)\bigr)^{\top}\xi\bigl(e^{-(Z^{0} (h))^{\top}\xi}-1\bigr)F(d\xi)\right)^{2} e^{\lambda s} \,ds \leq CM_{0}^{4}K_{0}^{2}. \end{aligned}$$

In view of the form of \(\partial_{s} \kappa^{i}_{3}(h)\) and \(\partial_{s} \kappa _{5}(h)\) as given in (4.11), (4.12), this implies that \(\kappa_{j}^{i}(H^{\lambda}_{m+1}) \subseteq H^{\lambda,0}_{1}\) for \(j=3,5\). We have thus shown that

$$\kappa^{i}(H^{\lambda}_{m+1}) \subseteq H^{\lambda,0}_{1}. $$

For \(h_{1}, h_{2} \in H^{\lambda}_{m+1}\) we obtain

$$\begin{aligned} \| \kappa^{i}_{1}(h_{1})-\kappa^{i}_{1}(h_{2})\|_{\lambda,1}&\leq C(L_{0}+L_{i})\| h_{1}-h_{2}\|_{\lambda, m+1},\\ \| \kappa^{i}_{2}(h_{1})-\kappa^{i}_{2}(h_{2})\|_{\lambda,1}& \leq C(2M_{i}+2M_{0})(L_{0}+L_{i})\|h_{1}-h_{2}\|_{\lambda, m+1},\\ \| \kappa_{4}(h_{1})-\kappa_{4}(h_{2})\|_{\lambda,1}& \leq C2M_{0}L_{0}\|h_{1}-h_{2}\|_{\lambda, m+1}. \end{aligned}$$

Furthermore, due to (4.11) and (4.12), we can estimate

$$\begin{aligned} \| \kappa^{i}_{3}(h_{1})-\kappa^{i}_{3}(h_{2})\|_{\lambda,1}^{2}& \leq4 (I^{i}_{1}+I^{i}_{2}+I^{i}_{3}+I^{i}_{4}),\\ \| \kappa_{5}(h_{1})-\kappa_{5}(h_{2})\|_{\lambda,1}^{2}& \leq4 (J_{1}+J_{2}+J_{3}+J_{4}), \end{aligned}$$

where

$$\begin{aligned} I^{i}_{1} &=\int_{\mathbb{R}_{+}}\bigg(\int\Big(\big(\zeta^{i0}(h_{1})(s)\big)^{\top }\xi\Big)^{2}e^{u_{i}^{\top} \hat{\xi}}\\ &\phantom{I^{i}_{1}=\int_{\mathbb{R}_{+}}\bigg(\int}{}\times\bigl(e^{(Z^{i0}(h_{1})(s))^{\top}\xi} -e^{(Z^{i0}(h_{2})(s))^{\top}\xi}\bigr)K^{\hat{Y},X}(d\hat{\xi}, d\xi)\bigg)^{2} e^{\lambda s} \,ds, \\ I^{i}_{2}&=\int_{\mathbb{R}_{+}}\bigg(\int e^{u_{i}^{\top} \hat{\xi}+ (Z^{i0}(h_{2})(s))^{\top}\xi}\bigg(\Big(\big(\zeta^{i0}(h_{1})(s)\big)^{\top }\xi\Big)^{2}\\ &\phantom{I^{i}_{2}=\int_{\mathbb{R}_{+}}\bigg(\int e^{u_{i}^{\top} \hat {\xi}+ (Z^{i0}(h_{2})(s))^{\top}\xi}}{} -\Big(\big(\zeta^{i0}(h_{2})(s)\big)^{\top}\xi\Big)^{2}\bigg) K^{\hat{Y},X}(d\hat{\xi}, d\xi)\bigg)^{2}e^{\lambda s} \,ds, \\ I^{i}_{3}&=\int_{\mathbb{R}_{+}}\bigg(\int\frac{d}{ds}\bigl(\zeta ^{i0}(h_{1})(s)\bigr)^{\top}\xi e^{u_{i}^{\top} \hat{\xi}}\\ &\phantom{I^{i}_{3}=\int_{\mathbb{R}_{+}}\bigg(\int}{}\times\bigl(e^{(Z^{i0}(h_{1})(s))^{\top}\xi} -e^{(Z^{i0}(h_{2})(s))^{\top}\xi}\bigr)K^{\hat{Y},X}(d\hat{\xi}, d\xi)\bigg)e^{\lambda s} \,ds,\\ I^{i}_{4}&=\int_{\mathbb{R}_{+}}\bigg(\int\Bigl(e^{u_{i}^{\top} \hat{\xi}+ (Z^{i0}(h_{2})(s))^{\top}\xi}-1\Bigr) \bigg(\frac{d}{ds}\bigl(\zeta ^{i0}(h_{1})(s)\bigr)^{\top}\xi\\ &\phantom{I^{i}_{4}=\int_{\mathbb{R}_{+}}\bigg(\int}{} -\frac{d}{ds}\big(\zeta^{i0}(h_{2})(s)\big)^{\top}\xi\bigg)K^{\hat{Y} ,X}(d\hat{\xi}, d\xi)\bigg)^{2}e^{\lambda s} \,ds, \\ J_{1}&=\int_{\mathbb{R}_{+}}\bigg(\int\Big(\big(\zeta^{0}(h_{1})(s)\big)^{\top }\Big)^{2} \\ &\phantom{J_{1}=\int_{\mathbb{R}_{+}}\bigg(}{}\times\bigl(e^{-(Z^{0} (h_{1})(s))^{\top}\xi} -e^{-(Z^{0} (h_{2})(s))^{\top}\xi}\big)F(d\xi)\bigg)^{2}e^{\lambda s} \,ds,\\ J_{2}&=\int_{\mathbb{R}_{+}}\bigg(\int e^{-(Z^{0} (h_{2})(s))^{\top}\xi}\\ &\phantom{J_{2}=\int_{\mathbb{R}_{+}}\bigg(}{}\times\bigg(\Big(\big(\zeta ^{0}(h_{1})(s)\big)^{\top}\xi\Big)^{2} -\Big(\big(\zeta^{0}(h_{2})(s)\big)^{\top}\xi\Big)^{2}\bigg)F(d\xi )\bigg)^{2} e^{\lambda s} \,ds,\\ J_{3}&=\int_{\mathbb{R}_{+}}\bigg(\int\frac{d}{ds}\big(\zeta ^{0}(h_{1})(s)\big)^{\top}\xi\\ &\phantom{J_{3}=\int_{\mathbb{R}_{+}}\bigg(}{}\times\big(e^{-(Z^{0} (h_{1})(s))^{\top}\xi} -e^{-(Z^{0} (h_{2})(s))^{\top}\xi}\big)F(d\xi)\bigg)^{2}e^{\lambda s} \,ds,\\ J_{4}&=\int_{\mathbb{R}_{+}}\bigg(\int\big(e^{-(Z^{0} (h_{2})(s))^{\top }\xi }-1\big)\\ &\phantom{J_{4}=\int_{\mathbb{R}_{+}}\bigg(}{}\times\bigg(\frac {d}{ds}\big(\zeta^{0}(h_{1})(s)\big)^{\top}\xi -\frac{d}{ds}\big(\zeta^{0}(h_{2})(s)\big)^{\top}\xi\bigg)F(d\xi )\bigg)^{2}e^{\lambda s} \,ds. \end{aligned}$$

We get for all \(\xi\in\mathbb{R}^{d}\), \(s \in\mathbb{R}_{+}\) that

$$\begin{aligned} \bigl|e^{(Z^{i0}(h_{1})(s))^{\top}\xi}-e^{(Z^{i0}(h_{2}(s)))^{\top}\xi }\bigr| \leq& Ce^{(C_{0}+C_{i})\| \xi\|_{d}}\bigl( \|\zeta^{i}(h_{1})-\zeta^{i}(h_{2})\|_{\lambda,d}\bigr. \\ &\bigl.{}+\|\zeta^{0}(h_{1})-\zeta^{0}(h_{2})\|_{\lambda,d}\bigr)\| \xi \| _{d},\\ \bigl|e^{-(Z^{0} (h_{1}))^{\top}\xi}-e^{-(Z^{0} (h_{2}))^{\top}\xi}\bigr| \leq& Ce^{C_{0}\| \xi\|_{d}}\|\zeta^{0}(h_{1})-\zeta^{0}(h_{2})\|_{\lambda ,d}\| \xi\|_{d}. \end{aligned}$$

Therefore,

$$\begin{aligned} I^{i}_{1}&\leq C\int_{\mathbb{R}_{+}}\bigg(\int\Big(\big(\zeta ^{i0}(h_{1})(s)\big)^{\top}\xi\Big)^{2}e^{\|u_{i}\|_{n} \|\hat{\xi}\| _{n}+(C_{0}+C_{i})\| \xi\|_{d}}\\ &\quad{}\times\bigl( \|\zeta^{i}(h_{1})-\zeta^{i}(h_{2})\|_{\delta,d}+\|\zeta^{0}(h_{1})-\zeta ^{0}(h_{2})\|_{\lambda,d}\bigr)\| \xi\|_{d} K^{\hat{Y},X}(d\hat{\xi}, d\xi)\bigg)^{2} e^{\lambda s} \,ds\\ & \leq CK^{2}_{i}(M_{i}^{4}+M_{0}^{4})(L^{2}_{i}+L_{0}^{2})\|h_{1}-h_{2}\|^{2}_{\lambda,m+1},\\ J_{1}&\leq C\int_{\mathbb{R}_{+}}\bigg(\int \Big(\big(\zeta ^{0}(h_{1})(s)\big)^{\top}\xi\Big)^{2} e^{C_{0}\| \xi\|_{d}}\|\zeta ^{0}(h_{1})-\zeta ^{0}(h_{2})\|_{\lambda,d}\| \xi\|_{d}F(d\xi)\bigg)^{2}e^{\lambda s} \,ds\\ &\leq C K^{2}_{0}M_{0}^{4}L_{0}^{2}\|h_{1}-h_{2}\|^{2}_{\lambda,m+1}. \end{aligned}$$

Moreover, for every \(s \in\mathbb{R}_{+}\), we obtain

$$\begin{aligned} &\int e^{\|u_{i}\|_{n} \|\hat{\xi}\|_{n}+ (C_{0}+C_{i})\|\xi\|_{d}} \Big(\big(\zeta^{i0}(h_{1})(s)\big)^{\top}\xi+\big(\zeta ^{i0}(h_{2})(s)\big)^{\top}\xi\Big)^{2} K^{\hat{Y},X}(d\hat{\xi}, d\xi) \\ &\quad\leq2C(M_{i}^{2}+M_{0}^{2})K_{i},\\ &\int e^{C_{0}\|\xi\|_{d}} \Big(\big(\zeta^{0}(h_{1})(s)\big)^{\top}\xi +\big(\zeta^{0}(h_{2})(s)\big)^{\top}\xi\Big)^{2}F(d\xi)\leq2CM_{0}^{2}K_{0}. \end{aligned}$$

Hence,

$$\begin{aligned} I^{i}_{2}&\leq2C(M_{i}+M_{0})K_{i}\int e^{\|u_{i}\|_{n} \|\hat{\xi}\|_{n}+ (C_{0}+C_{i})\|\xi \|_{d}}\\ &\quad{} \times\int_{\mathbb{R}_{+}}\Big(\bigl(\zeta ^{i}(h_{1})(s)-\zeta^{i}(h_{2})(s)\bigr)^{\top}\xi\\ &\phantom{=:\times\int_{\mathbb{R}_{+}}\Big(}{}+\bigl(\zeta ^{0}(h_{2})(s)-\zeta^{0}(h_{1})(s)\bigr)^{\top}\xi\Big)^{2}e^{\lambda s} ds K^{\hat{Y},X}(d\hat{\xi}, d\xi)\\ &\leq2C(M_{i}+M_{0})K_{i}^{2} (L_{0}^{2}+L_{i}^{2})\|h_{1}-h_{2}\|^{2}_{\lambda,m+1},\\ J_{2} &\leq2CM_{0}K_{0}\int e^{C_{0}\|\xi\|_{d}}\int_{\mathbb{R}_{+}}\left (\bigl(\zeta^{0}(h_{2})(s)-\zeta^{0}(h_{1})(s)\bigr)^{\top}\xi\right )^{2}e^{\lambda s} ds F(d\xi)\\ &\leq2CM_{0}K_{0}^{2} L_{0}^{2}\|h_{1}-h_{2}\|^{2}_{\lambda,m+1}. \end{aligned}$$

Moreover,

$$\begin{aligned} I^{i}_{3}&\leq CK_{i}(L^{2}_{i}+L_{0}^{2})\|h_{1}-h_{2}\|^{2}_{\lambda,m+1}\int e^{\|u_{i}\| _{n} \|\hat{\xi}\|_{n}+(C_{0}+C_{i})\| \xi\|_{d}} \\ &\quad{}\times\int_{\mathbb{R}_{+}} \left(\frac{d}{ds}\bigl(\zeta ^{i0}(h_{1})(s)\bigr)^{\top}\xi\right)^{2} e^{\lambda s} \,ds K^{\hat{Y},X}(d\hat{\xi}, d\xi)\\ &\leq CK^{2}_{i}(L^{2}_{i}+L_{0}^{2})(M_{0}+M_{i})\|h_{1}-h_{2}\|^{2}_{\lambda,m+1},\\ J_{3} &\leq CK_{0}L_{0}^{2}\|h_{1}-h_{2}\|^{2}_{\lambda,m+1}\int e^{C_{0}\|\xi\| _{d}}\int _{\mathbb{R}_{+}}\left(\frac{d}{ds}\bigl(\zeta^{0}(h_{1})(s)\bigr)^{\top}\xi \right)^{2} e^{\lambda s} \,ds F(d\xi)\\ &\leq CK^{2}_{0}L_{0}^{2}M_{0}\|h_{1}-h_{2}\|^{2}_{\lambda,m+1}. \end{aligned}$$

Finally, we have

$$\begin{aligned} I^{i}_{4}&\leq C\int_{\mathbb{R}_{+}}\bigg(\int e^{\|u_{i}\|_{n}\|\hat{\xi}\| _{n}+(C_{0}+C_{i}) \|\xi\|_{d}}\bigl(\|u_{i}\|_{n}\|\hat{\xi}\|_{n}+\| \zeta ^{i0}(h)\| _{\lambda,d}\|\xi\|_{d}\bigr)\\ &\phantom{\leq C\int_{\mathbb{R}_{+}}\bigg(}{} \times\left(\frac {d}{ds}\bigl(\zeta^{i0}(h_{1})(s)\bigr)^{\top}\xi-\frac{d}{ds}\bigl(\zeta ^{i0}(h_{2})(s)\bigr)^{\top}\xi\right)K^{\hat{Y},X}(d\hat{\xi}, d\xi)\bigg)^{2}e^{\lambda s} \,ds\\ &\leq C(u_{i}+M_{0}+M_{i})^{2}K_{i}^{2}(L_{0}+L_{i})^{2}\|h_{1}-h_{2}\|^{2}_{\lambda ,m+1},\\ J_{4}&\leq C\int_{\mathbb{R}_{+}}\bigg(\int e^{C_{0} \|\xi\|_{d}}\| \zeta ^{0}(h)\| _{\lambda,d}\| \xi\|_{d} \\ &\phantom{\leq C\int_{\mathbb{R}_{+}}\bigg(}{}\times\bigg(\frac {d}{ds}\bigl(\zeta^{0}(h_{1})(s)\bigr)^{\top}\xi -\frac{d}{ds}\bigl(\zeta^{0}(h_{2})(s)\bigr)^{\top}\xi\bigg)F(d\xi )\bigg)^{2}e^{\lambda s} \,ds,\\ &\leq CM_{0}^{2}K_{0}^{2}L_{0}^{2}\|h_{1}-h_{2}\|^{2}_{\lambda,m+1}. \end{aligned}$$

Summing up, we have shown that there exist constants \(Q_{i}>0\) such that condition (4.13) is satisfied for all \(h_{1},h_{2} \in H_{m+1}^{\lambda}\). □

Proof of Proposition 4.10

For every \((\omega,\omega')\in\widetilde{\varOmega}\) and \(t\geq0\), define the process \(Y^{\perp}_{t}(\omega,\omega') := y_{0} + J_{t}(\omega')\), for some starting value \(y_{0}\in\mathbb{R}_{+}\). Clearly, \(Y^{\perp}\) is a pure jump \((\mathcal{G}_{t})\)-adapted process. In order to prove the existence of a probability measure \(\widetilde {\mathbb{Q}}\) such that the jump measure of \(Y^{\perp}\) with respect to the two filtrations \((\mathcal{G}_{t})_{t\geq0}\) and \((\overline {\mathcal{G}}_{t})_{t\geq 0}\) is given by \(K_{t}(\omega, Y^{\perp}_{t-}(\omega,\omega'),d\xi)dt\) and \(\widetilde{\mathbb{Q}}|_{\mathcal{F}}=\mathbb{Q}\) holds true, we rely on [34, Theorem 3.6].

For all \(\omega\in\varOmega\), \(y\in\mathbb{R}_{+}\) and \(t\geq0\), let us first extend the definition of \(K_{t}(\omega,y, d\xi)\) as in (4.15), (4.16) to \(y\in\mathbb{R}_{-}\) by requiring that it is supported on \([-|y|, \infty)\) and by setting \(p_{t}(y,\omega )=p_{t}(-y, \omega)\) for \(y\in\mathbb{R}_{-}\). Due to Assumption 4.8, the measure \(K_{t}(\omega,Y^{\perp}_{t-}(\omega ,\omega'), d\xi)dt\) defined via the moment problem (4.15), (4.16) is a positive random measure on \(\mathbb{R}_{+} \times\mathbb{R}\). Since \(Y^{\perp}\) is càdlàg and \((\mathcal{G}_{t})\)-adapted and since \((p_{t}(\cdot,y))_{t\geq0}\) is \((\mathcal{F}_{t})\)-predictable and depends in a measurable way on \(y\), the process \(p_{t}(\omega, Y^{\perp}_{t-}(\omega ,\omega'))\) is \((\mathcal{G}_{t})\)-predictable and the same \((\mathcal{G} _{t})\)-predictability, and hence \((\overline{\mathcal {G}}_{t})\)-predictability, is inherited by \(K_{t}(\omega,Y^{\perp}_{t-}(\omega,\omega'), d\xi)\). Let us then define the \((\overline{\mathcal{G}}_{t})\)-predictable random measure \(\nu\) by

$$\nu(\widetilde{\omega},dt,d\xi)=\left\{ \textstyle\begin{array}{l@{\quad}l} K_{t}\bigl(\omega, Y^{\perp}_{t-}(\omega,\omega'),d\xi\bigr)dt, & t < T_{\infty},\\ 0, & t \geq T_{\infty}. \end{array}\displaystyle \right. $$

[34, Theorem 3.6] implies that there exists a unique probability kernel ℙ from \((\varOmega,\mathcal{F})\) to ℋ such that \(\nu\) is the \((\overline{\mathcal{G}}_{t})\)-compensator of the random measure \(\mu\) associated with the jumps of \(J\). On \((\widetilde {\varOmega}, \mathcal{G})\), we then define the probability measure \(\widetilde{\mathbb{Q}}\) by setting \(\widetilde{\mathbb {Q}}(d\widetilde{\omega})=\mathbb{Q}(d\omega)\mathbb{P}(\omega ,d\omega')\), whose restriction to ℱ is equal to ℚ. Moreover, since \(Y^{\perp}\) is \((\mathcal{G}_{t})\)-adapted and \(K_{t}(\omega ,Y^{\perp}_{t-}(\omega,\omega'), d\xi)\) is \((\mathcal {G}_{t})\)-predictable, the random measure \(\nu\) is also the \((\mathcal{G}_{t})\)-compensator of the jump measure of \(J\). Since for every \(\omega\in\varOmega\), \(y\in\mathbb{R}_{+}\) and \(t\geq 0\), the measure \(K_{t}(\omega,y,d\xi)\) is supported by \([-y,\infty)\), the process \(Y^{\perp}\) takes values in \(\mathbb{R}_{+}\).

It remains to show that \(Y^{\perp}\) is of finite activity (equivalently, that \(T_{\infty}=\infty\) \(\widetilde{\mathbb{Q}}\)-a.s.). Since \(g_{m+1}(\xi)\geq1\) for all \(\xi\in\mathbb{R}\) and due to condition (4.16), it holds that

$$\begin{aligned} \mathbb{E}^{\widetilde{\mathbb{Q}}}\bigl[\mu([0,T]\times\mathbb {R})\bigr] &=\mathbb{E}^{\widetilde{\mathbb{Q}}}\bigl[\nu([0,T]\times\mathbb {R})\bigr] =\mathbb{E}^{\widetilde{\mathbb{Q}}}\left[\int_{0}^{T} K_{t}\bigl(\omega, Y_{t-}(\omega,\omega'), \mathbb{R}\bigr)\,dt\right ]\\ &\leq\mathbb{E}^{\widetilde{\mathbb{Q}}}\left[\int_{0}^{T}\int g_{m+1}(\xi)K_{t}\bigl(\omega, Y_{t-}(\omega,\omega'),d\xi\bigr)\,dt\right] \\ &= \mathbb{E}^{\widetilde{\mathbb{Q}}}\left[\int_{0}^{T}p^{m+1}_{t}\bigl(\omega,Y_{t-}(\omega ,\omega')\bigr)\,dt\right] \leq\bar{H}T, \end{aligned}$$

due to the uniform boundedness of the processes \(\{(p^{m+1}_{t}(\cdot ,y))_{t\geq0};y\in\mathbb{R}_{+}\}\). This implies that \(\mu([0,T]\times\mathbb{R})<\infty\) \(\widetilde {\mathbb{Q}}\)-a.s. for each \(T \geq0\), and as a consequence, \(\widetilde{\mathbb{Q}}[T_{\infty}<\infty]=0\). □

Proof of Lemma 4.11

(i) For \(t\geq0\), let \(H\) be a bounded \(\mathcal{H}_{t}\)-measurable random variable, \(F\) a bounded \(\mathcal{F}_{t}\)-measurable random variable and \(A\in\mathcal{F}_{\infty}\). As can be deduced from the proof of Proposition 4.10, the \(\mathcal{F}_{\infty }\)-conditional law of \((\omega'(s))_{s\in[0,t]}\) under \(\widetilde {\mathbb{Q}}\) is \(\mathcal{F}_{t}\)-measurable (compare also with [20, part (iv) of Theorem 5.1]). In particular, this means that \(\mathbb{E}^{\widetilde{\mathbb {Q}}}[H|\mathcal{F}_{\infty}]=\mathbb{E} ^{\widetilde{\mathbb{Q}}}[H|\mathcal{F}_{t}]\). In turn, this implies that

$$\begin{aligned} \mathbb{E}^{\widetilde{\mathbb{Q}}}[FH\textbf{1}_{A}] &= \mathbb{E}^{\widetilde{\mathbb{Q}}}\big[F\mathbb{E}^{\widetilde {\mathbb{Q}}}[H|\mathcal{F}_{\infty}]\textbf{1}_{A}\big]\\ &= \mathbb{E}^{\widetilde{\mathbb{Q}}}\big[F\mathbb{E}^{\widetilde {\mathbb{Q}}}[H|\mathcal{F}_{t}]\textbf{1}_{A}\big] = \mathbb{E}^{\widetilde{\mathbb{Q}}}\big[\mathbb{E}^{\widetilde {\mathbb{Q}}}[FH|\mathcal{F}_{t}]\textbf{1}_{A}\big]. \end{aligned}$$

By a monotone class argument, this means that \(\mathbb{E}^{\widetilde {\mathbb{Q}}}[G|\mathcal{F} _{\infty}]=\mathbb{E}^{\widetilde{\mathbb{Q}}}[G|\mathcal{F}_{t}]\) for every bounded \(\mathcal{G} _{t}\)-measurable random variable \(G\). It is well known (see e.g. [38, Proposition 5.9.1.1]) that the latter property is equivalent to the fact that all \((\widetilde {\mathbb{Q}},(\mathcal{F} _{t})_{t\geq0})\)-martingales are also \((\widetilde{\mathbb{Q}},(\mathcal{G}_{t})_{t\geq 0})\) martingales. Due to the fact that \(\widetilde{\mathbb{Q}}|_{\mathcal{F}_{\infty}}=\mathbb{Q}\), this implies that all \((\mathbb{Q},(\mathcal{F} _{t})_{t\geq0})\)-local martingales are also \((\widetilde{\mathbb {Q}},(\mathcal{G}_{t})_{t\geq 0})\)-local martingales. As a consequence, every \((\mathbb{Q},(\mathcal {F}_{t})_{t\geq 0})\)-semimartingale is also a \((\widetilde{\mathbb{Q}},(\mathcal {G}_{t})_{t\geq0})\) semimartingale. Moreover, since semimartingale characteristics can be characterized in terms of local martingales (see e.g. [36, Theorem II.2.21]), this implies that \((X,\hat{Y})\) is a semimartingale with respect to \((\widetilde{\mathbb{Q}},(\mathcal{G}_{t})_{t\geq0})\) with unchanged characteristics.

(ii) Since \(Y^{\perp}\) is a pure jump process, in order to prove its local independence with respect to \((X,\hat{Y})\), it suffices to show that \(Y^{\perp}\) and \((X,\hat{Y})\) never jump together. In view of (C.1), this reduces to showing that

$$\widetilde{\mathbb{Q}}[\exists t>0 \mid\varDelta Y^{\perp}_{t}\neq 0\text{ and }\varDelta X_{t}\neq0]=0. $$

Let \(\mathfrak{T}\) be the set of jump times of \(X\). Since \(X\) is càdlàg, the set \(\mathfrak{T}\) is countable (see e.g. [36, Proposition I.1.32]) and similarly as in [39, Theorem 4.7],

$$\begin{aligned} \widetilde{\mathbb{Q}}[\exists t>0 \mid\varDelta Y^{\perp}_{t}\neq 0\text{ and }\varDelta X_{t}\neq0] &\leq\mathbb{E}^{\widetilde{\mathbb{Q}}}\left[\sum_{t\in \mathfrak{T}}\textbf{1}_{\{\varDelta Y_{t}^{\perp}\neq0\}}\right] \\ &= \mathbb{E}^{\widetilde{\mathbb{Q}}}\left[\sum_{t\in\mathfrak {T}}\mathbb{E}^{\widetilde{\mathbb{Q}}}[\textbf{1}_{\{ \varDelta Y_{t}^{\perp}\neq0\}}|\mathcal{F}_{\infty}]\right]=0, \end{aligned}$$

where the last equality follows from the fact that \(\mathbb {E}^{\widetilde{\mathbb{Q}}}[\textbf{1} _{\{\varDelta Y_{t}^{\perp}\neq0\}}|\mathcal{F}_{\infty}]=0\) for all \(t>0\) since due to Proposition 4.10, the jump measure of \(Y^{\perp}\) with respect to the larger filtration \((\overline{\mathcal {G}}_{t})_{t\geq0}\) (which satisfies \(\overline{\mathcal{G}}_{0}=\mathcal{F}_{\infty }\otimes\{\emptyset,\varOmega '\}\)) is absolutely continuous with respect to Lebesgue measure, so that \(Y^{\perp}\) does not have any fixed time of discontinuity (see e.g. [36, Lemma II.2.54]).

In order to prove the last assertion, note that condition (4.16) implies that condition \(I(0,1)\) from [40] is satisfied, since for all \(y\in\mathbb{R}_{+}\), \(T\geq0\) and \(i=1,\ldots,m\),

$$\begin{aligned} &\sup_{t\in[0,T]}\mathbb{E}^{\widetilde{\mathbb{Q}}}\left[\exp \left(\int_{0}^{t}\int\bigl(e^{u_{i}\xi}(u_{i}\xi-1)+1\bigr)K_{s}\bigl(\omega,Y^{\perp }_{s-}(\omega ,\omega'),d\xi\bigr)\,ds\right)\right] \\ &\quad\leq \sup_{t\in[0,T]}\mathbb{E}^{\widetilde{\mathbb{Q}}}\left[\exp \left((1+u_{m})\int_{0}^{t}\int g_{m+1}(\xi)K_{s}\bigl(\omega,Y^{\perp}_{s-}(\omega,\omega'),d\xi \bigr)\,ds\right)\right] \\ &\quad= \sup_{t\in[0,T]}\mathbb{E}^{\widetilde{\mathbb{Q}}}\left[\exp \left((1+u_{m})\int _{0}^{t}p^{m+1}_{s}\bigl(\omega,Y^{\perp}_{s-}(\omega,\omega')\bigr) \,dt\right)\right] \leq e^{(1+u_{m})T\bar{H}}< \infty. \end{aligned}$$

Moreover, condition (4.16) can be easily shown to imply that

$$\int_{0}^{T}\int|\xi e^{u_{i}\xi}-\xi|K_{t}\big(\omega,Y^{\perp }_{t-}(\omega ,\omega'),d\xi\big)\,dt< \infty \quad\widetilde{\mathbb{Q}}\text{-a.s.} $$

for all \(T\geq0\). Hence, [40, Theorem 3.2] implies that

$$\bigg(\exp\Big(u_{i}Y^{\perp}_{t}-\int_{0}^{t}\varPsi^{Y^{\perp }}_{s}(u_{i})\,ds\Big)\bigg)_{t\in[0,T]} $$

is a uniformly integrable \((\widetilde{\mathbb{Q}},(\overline {\mathcal{G}}_{t})_{t\geq 0})\)-martingale, for all \(i=1,\ldots,m\). In turn, because \(T\geq0\) is arbitrary, this proves the \((\widetilde{\mathbb{Q}},(\overline {\mathcal{G}}_{t})_{t\geq 0})\)-martingale property of \((\exp(u_{i}Y^{\perp}_{t}-\int _{0}^{t}\varPsi ^{Y^{\perp}}_{s}(u_{i})\,ds))_{t\geq0}\), for all \(i=1,\ldots,m\). Finally, since the latter process is \((\mathcal{G}_{t})\)-adapted, it is also a martingale in the smaller filtration \((\mathcal{G}_{t})_{t\geq0}\). □

Proof of Theorem 4.12

Due to Lemma 4.11, the local exponent of \(\hat{Y}\) with respect to the extended filtered probability space \((\varOmega,\mathcal {G},(\mathcal{G} _{t})_{t\geq0},\widetilde{\mathbb{Q}})\) is still given by \(\varPsi^{\hat {Y}}\) and it holds that \(Y^{\parallel}=(\hat{Y})^{\parallel}=\hat{Y}\). Since \(Y^{\perp}\) and \(Y-Y^{\perp}=\hat{Y}\) are locally independent (see Lemma 4.11), the consistency condition (3.11) directly follows from condition (4.15).

In order to prove the martingale property of the process in (3.10), note first that the \((\widetilde{\mathbb {Q}},(\mathcal{G}_{t})_{t\geq 0})\)-martingale property of \((\exp(u_{i}Y^{\perp}_{t}-\int _{0}^{t}\varPsi ^{Y^{\perp}}_{s}(u_{i})ds))_{t\geq0}\) (see Lemma 4.11), condition (iv) in Definition 4.1 and the local independence of \(Y^{\perp}\) and \((X,\hat{Y})\) on \((\varOmega,\mathcal{G},(\mathcal{G}_{t})_{t\geq0},\widetilde{\mathbb {Q}})\) imply that the process in (3.10) is a \((\widetilde{\mathbb{Q}},(\mathcal {G}_{t})_{t\geq0})\)-local martingale, for all \(i=1,\ldots,m\). Being a nonnegative local martingale, it is also a supermartingale by Fatou’s lemma. Hence, to establish the true martingale property, it suffices to observe that since \(\overline{\mathcal{G}}_{0}=\mathcal {F}_{\infty}\otimes\{ \emptyset,\varOmega'\}\), it holds for all \(T\geq0\) and \(i=1,\ldots,m\) that

$$\begin{aligned} &\mathbb{E}^{\widetilde{\mathbb{Q}}}\bigg[\exp\bigg(u_{i}Y_{T}+\int _{0}^{T}\bigl(\varSigma ^{i}_{s}(T)-\widetilde{\varSigma}_{s}(T)\bigr)dX_{s} \\ &\phantom{\mathbb{E}^{\widetilde{\mathbb{Q}}}\bigg[\exp\bigg(}{} -\int_{0}^{T}\varPsi^{Y,X}_{s}\Big(\bigl(u_{i},\varSigma^{i\top }_{s}(T)-\widetilde {\varSigma}^{\top}_{s}(T)\bigr)^{\top}\Big)\,ds\bigg)\bigg] \\ &= \mathbb{E}^{\widetilde{\mathbb{Q}}}\Bigg[\exp\bigg(u_{i}\hat {Y}_{T}+\int_{0}^{T}\bigl(\varSigma ^{i}_{s}(T)-\widetilde{\varSigma}_{s}(T)\bigr)\,dX_{s} \\ &\phantom{= \mathbb{E}^{\widetilde{\mathbb{Q}}}\Bigg[\exp\bigg(}{} -\int_{0}^{T}\varPsi^{\hat{Y},X}_{s}\Big(\bigl(u_{i},\varSigma^{i\top }_{s}(T)-\widetilde {\varSigma}^{\top}_{s}(T)\bigr)^{\top}\Big)\,ds\bigg) \\ &\phantom{= \mathbb{E}^{\widetilde{\mathbb{Q}}}\Bigg[}{}\times \mathbb{E}^{\widetilde{\mathbb{Q}}}\bigg[\exp\bigg(u_{i}Y_{T}^{\perp}-\int_{0}^{T}\varPsi^{Y^{\perp}}_{s}(u_{i})\,ds\bigg)\bigg|\overline {\mathcal{G}}_{0}\bigg]\Bigg] \\ &= \mathbb{E}^{\widetilde{\mathbb{Q}}}\bigg[\exp\bigg(u_{i}\hat {Y}_{T}+\int_{0}^{T}\bigl(\varSigma ^{i}_{s}(T)-\widetilde{\varSigma}_{s}(T)\bigr)\,dX_{s}\\ &\phantom{= \mathbb{E}^{\widetilde{\mathbb{Q}}}\bigg[\exp\bigg(}{} -\int_{0}^{T}\varPsi^{\hat{Y},X}_{s}\Big(\bigl(u_{i},\varSigma^{i\top }_{s}(T)-\widetilde {\varSigma}^{\top}_{s}(T)\bigr)^{\top}\Big)\,ds\bigg)\bigg]e^{u_{i}Y^{\perp}_{0}} \\ &= \mathbb{E}^{\mathbb{Q}}[\exp(u_{i} Y_{0})], \end{aligned}$$

where in the last equality, we have used \(\widetilde{\mathbb {Q}}|_{\mathcal{F}}=\mathbb{Q}\) and the \((\mathbb{Q},(\mathcal{F}_{t})_{t\geq0})\)-martingale property of (4.1). □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cuchiero, C., Fontana, C. & Gnoatto, A. A general HJM framework for multiple yield curve modelling. Finance Stoch 20, 267–320 (2016). https://doi.org/10.1007/s00780-016-0291-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00780-016-0291-5

Keywords

Mathematics Subject Classification (2010)

JEL Classification

Navigation