Financial applications are given to illiquidity modeling using regime-switching time- inhomogeneous Levy price dynamics, to regime-switching Levy driven diffusion based price dynamics, and to a generalized version of the multi-asset model of price impact from distress selling, for which we retrieve and generalize their diffusion limit result for the price process.
6.1. Regime-Switching Inhomogeneous Lévy Based Stock Price Dynamics and Application to Illiquidity Modeling
For specific examples in this section, we state all the proofs related to regularity of the propagator, compact containment criterion, etc. It is worth noting that we focus below on the Banach space
, but that, in the article [
15], other
-type Banach spaces denoted by
are considered (in addition to
), for which a suitable weighted sup-norm is introduced (see their Lemma 2.6).
We first give a brief overview of the proposed model and the intuition behind it, and then we go into more formal details, presenting precisely all the various variables of interest, checking assumptions we have made up to this point, etc. We consider a regime-switching inhomogeneous Lévy-based stock price model, very similar in the spirit to the recent article [
4]. In short, an inhomogeneous Lévy process differs from a classical Lévy process in the sense that it has time-dependent (and absolutely continuous) characteristics. We let
be a collection of such
-valued inhomogeneous Lévy processes with characteristics
, and we define:
for some bounded function α. We give below a financial interpretation of this function α, as well as reasons we consider a regime-switching model. In this setting,
f represents a contingent claim on a (
d-dimensional) risky asset
S having regime-switching inhomogeneous Lévy dynamics driven by the processes
: on each random time interval
, the risky asset is driven by the process
. Indeed, we have the following representation, for
(to make clear that the expectation below is taken with respect to
ω embedded in the process
L and not
):
where we have denoted for clarity:
The random evolution
represents in this case the present value of the contingent claim
f of maturity
t on the risky asset
S, conditionally on the regime switching process
: indeed, remember that
is random, and that its randomness (only) comes from the Markov renewal process. Our main results in Theorems 8 and 11 allow approximating the impact of the regime-switching on the present value
of the contingent claim. Indeed, we get the following normal approximation, for small
ϵ:
The above approximation allows quantifying the risk inherent to regime-switchings occurring at a high frequency governed by ϵ. The parameter ϵ reflects the frequency of the regime-switchings and can therefore be calibrated to market data by the risk manager. For market practitioners, because of the computational cost, it is often convenient to have asymptotic formulas that allow them to approximate the present value of a given derivative, and by extent the value of their whole portfolio. In addition, the asymptotic normal form of the regime-switching cost allows the risk manager to derive approximate confidence intervals for his portfolio, as well as other quantities of interest such as reserve policies linked to a given model.
In the recent article [
4], an application to illiquidity is presented: the authors interpreted the state process
as a “
proxy for market liquidity with states (or regimes) representing the level of liquidity activity”. In their context, however,
is taken to be a continuous-time Markov chain, which is a specific case of semi-Markov process with exponentially distributed sojourn times
. With creative choices of the Markov renewal kernel, one could choose alternative distributions for the “waiting times” between regimes, thus generalizing the setting of [
4]. In [
4], the equivalent of our quantities
are the constants
, which represent an impact on the stock price at each regime-switching time. The authors stated: “
Typically, there is a drop of the stock price after a liquidity breakdown”. This justifies the use of the operators
D.
In addition to the liquidity modeling of [
4], another application of our framework could be to view the different states as a reflection of uncertainty on market data induced by, for example, a significant bid-ask spread on the observed market data. The different characteristics
would, in this context, reflect this uncertainty induced by the observables. The random times
would then correspond to times at which the observed market data (e.g., call option prices, or interest rates) switch from one value to another. As these switching times are typically very small (∼ms), an asymptotic expansion as
would make perfect sense in order to get an approximation of the stock price at a scale ≫
, e.g., ∼min.
Now, let us formalize the model a bit more precisely. We let
,
for some
. Let
a probability space and let us begin with
a
-valued (time-inhomogeneous) Markov process on it with transition function
(
,
) and starting point
(in the sense of [
31], II.10.2). Formally,
,
,
p satisfies:
, is a probability measure.
is measurable.
.
(Chapman–Kolmogorov).
One example of such time-inhomogeneous Markov processes are additive processes (in the sense of [
31], I.1.6). This class of processes differs from Lévy processes only on the fact that increments are not assumed to be stationary. Additive processes are translation invariant (or spatially homogeneous) and we have (by [
31], II.10.4)
. We focus on additive processes in the following as they are rich enough to cover many financial applications. We define for
and
:
It can be proved that Γ is a regular
-contraction propagator and
we have
and
. The Lévy–Khintchine representation of such a process
(see [
32], 14.1) ensures that there exists unique
,
a family of
symmetric nonnegative-definite matrices and
a family of measures on
such that:
are called the spot characteristics of L. They satisfy the following regularity conditions:
, and
and : is symmetric nonnegative-definite and .
and as , , and such that for some .
If , , we can replace by in the Lévy–Khintchine representation of L, where . We denote by this other version of the spot characteristics of L.
Let
. We can consider a specific case of additive processes, called processes with independent increments and absolutely continuous characteristics (PIIAC), or inhomogeneous Lévy processes in [
33], for which there exists
,
a family of
symmetric nonnegative-definite matrices and
a family of measures on
satisfying:
where
denotes any norm on the space of
matrices.
are called the local characteristics of
L. We can notice by [
33] that PIIAC are semi-martingales, which is not the case of all additive processes. According to [
33], we have the following representation for
L:
where for
:
and
N the Poisson measure of
L (
is then called the compensated Poisson measure of
L).
is a
d-dimensional Brownian motion on
, independent from the jump part.
here stands for the unique symmetric nonnegative-definite square root of
. Sometimes it is convenient to write a Cholesky decomposition
and replace
by
in the previous representation (in this case, the Brownian motion
W would have to be replaced by another Brownian motion
, the point being that both
and
are processes with independent Gaussian increments with mean 0 and variance
). It can be shown (see [
15]) that the infinitesimal generator of the propagator Γ is given by:
and that
. If
is well-defined:
As a first example, we can consider the inhomogeneous Poisson process, for which
, local characteristics
, where the intensity function
s.t.
. We have, letting
:
As a second example, we can consider a risk process
L (used in insurance, for example):
where
is an inhomogeneous Poisson process,
is the premium intensity function and the
’s are iid random variables, independent from
N with common law
that represent the random payments that an insurance company has to pay. In this case, the local characteristics
. We get:
We can also consider a Brownian motion with time-dependent variance-covariance matrix
and drift
(local characteristics
):
To define the corresponding random evolution, we let
L be an inhomogeneous Lévy process taking value in
, independent of the semi-Markov process
. In this case, the ℝ-valued Lévy processes
are not necessarily independent but they have independent increments over non-overlapping time-intervals. Denote
the corresponding
-valued inhomogeneous Lévy processes, with local characteristics
. Define for
,
,
:
We define the jump operators, for
by:
where
, so that
and
. Assume that the ℝ-valued inhomogeneous Lévy processes
admit a second moment, namely they satisfy:
then it can be proved (see below) that the compact containment criterion is satisfied:
-CCC (Assumption 2). Assumption 1 will be satisfied provided the characteristics are smooth enough. Regarding Assumption 3, it is clear that
is the generator of the (inhomogeneous) Lévy propagator with “weighted” local characteristics
given by:
In particular, we notice that is indeed a Lévy measure on .
Proof that Γ is a propagator and that we have and . Γ satisfies the propagator equation because of the Chapman–Kolmogorov equation. Now, let us show that
we have
and
. Let
. Let us first start with
. Let
. By ([
31]), we get the representation
, where
is the distribution of
, i.e.,
for
. Let
and take any sequence
:
and denote
and
. By continuity of
f,
pointwise. Further,
. Therefore, by Lebesgue dominated convergence theorem, we get
. Therefore,
. By the same argument but now taking any sequence
:
, we get
and therefore
. Further, we get:
and therefore
. Let
,
(
). Take any sequence
:
,
and
. Then:
Let
and
. We have
pointwise since
. By the Mean Value theorem,
, and therefore
. Therefore, by Lebesgue dominated convergence theorem, we get:
Using the same argument as for , we get that since . Repeating this argument by computing successively every partial derivative up to order q by the relationship , we get . Further, the same way we got for , we get for : . Therefore, .
Proof that Γ is -strongly s-continuous, Y-strongly t-continuous. Let
,
and
:
Let
and
any sequence such that
. Let
and
. We have
by stochastic continuity of additive processes. By the Skorokhod’s representation theorem, there exists a probability space
and random variables
,
on it such that
,
and
. Let
, we therefore get:
Further, since
,
is uniformly continuous on
and
,
:
. Because
, for a.e.
,
. Therefore, we have that
. Further
. By Lebesgue dominated convergence theorem, we get:
We can notice that the proof strongly relies on the uniform continuity of f, and therefore on the topological properties of the space (which does not have). We prove that Γ is Y-strongly t-continuous exactly the same way, but now considering and any sequence such that , where if and if .
Proof that Γ is regular. By Taylor’s theorem, we get
,
,
:
observing that
by assumption. Therefore, by integrability assumption on the local characteristics and Theorem 3, we get the propagator evolution equations and therefore the regularity of Γ.
Proof that -CCC. We are going to apply Proposition 4. We showed
Vϵ is a
-contraction, so it remains to show the uniform convergence to 0 at infinity. We have the following representation, for
(to make clear that the expectation is with respect to
ω and not
):
where we have denoted for clarity
. In the following we drop the index
. Denote the
d components of the inhomogeneous Lévy process
by
(
). The multivariate version of Chebyshev’s inequality yields for any
d-vector of
integrable random variables
:
We apply this inequality to the d-vector having components
, where is the component of . If we denote by
the local characteristics of
, we have ([
32], proposition 3.13):
and so:
where
and
. Similarly, we have:
where
. Now, since
L is a
valued inhomogeneous Lévy process, the ℝ valued inhomogeneous Lévy processes
are not necessarily independent but they have independent increments over non-overlapping time intervals. This yields:
Overall, we get that for every
, there exists
such that:
On the other hand, there exists
:
. Define, as in the traffic case,
so that
. We have for
and
:
so that overall, for
and
(uniform in
,
ϵ) we have:
which completes the proof.
6.3. Multi-Asset Model of Price Impact from Distressed Selling: Diffusion Limit
Take, again,
. The setting of the recent article [
11] fits exactly into our framework. In this article, they consider a discrete-time stock price model for
d stocks
. It is assumed that a large fund
V holds
units of each asset
i. Denoting the
ith stock price and fund value at time
by, respectively,
and
, we have
, and the following dynamics are assumed for the stock prices:
where
is the constant time-step,
are i.i.d.
-valued centered random variables with covariance matrix
(i.e.,
=0,
);
is increasing, concave, equal to 0 on
and
, where
;
represents the depth of the market in asset
i, and
are constants. It is also assumed that:
so that the stock prices remain positive. The idea is that when the fund value
V reaches a certain “low” level
, market participants will start selling the stocks involved in that fund, inducing a correlation between the stocks, which “adds” to their fundamental correlations: this is captured by the function
g (which, if equal to 0, gives us the “standard” Black and Scholes setting).
The above setting is a particular case of our framework, where
for all
, and the operators
D have the following form:
where ∘ denotes composition of functions,
,
m is the vector with coordinates
,
has coordinate functions
defined by:
has coordinate functions
defined by:
and finally where we have extended the definition of
by the convention that
is the vector with coordinates
. We notice that the operator
D defined above does not depend on
x, and so we let
. By i.i.d. assumption on the random variables
, the process
is chosen such that it is a Markov chain with all rows of the transition matrix
P equal to each other, meaning that
is independent of
: in this context the ergodicity of the Markov chain is immediate, and we have
. It remains to choose the finite state space
: we assume that the random variables
can only take finitely many values, say
M, and we denote
these possible values. In the paper [
11], the authors considered that these random variables can have arbitrary distributions: here we approximate these distributions by finite state distributions. Note that in practice (e.g., on a computer), all distributions are in fact approximated by finite-state distributions, since the finite floating-point precision of the computer only allows finitely many values for a random variable. The parameter
M in fact plays no role in the analysis below. We let
. We have, for
and
:
In [
11], the times
are deterministic (so that
), but in our context we allow them to be possibly random, which can be seen for example as an “illiquidity extension” of the model in [
11], similar to what we have done in
Section 6.1. In addition, our very general framework allows one to be creative with the choices of the operators
D and Γ, leading to possibly interesting extensions of the model [
11] presented above. For example, one could simply “enlarge” the Markov chain
and consider the same model but with regime-switching parameters
and
, and one could carry on all the analysis below at almost no extra technical and computational cost. Since
, we let
, and we get:
where
, and:
Since
and
, it results that:
From the last expression, and remembering that each component
of
is centered, we get:
Let us denote the vector of initial log-spot prices
which elements are
(
), as well as the
d-dimensional price process
with elements:
In this context, the spot price changes at each jump time of the Markov renewal process
. As mentioned above, in [
11], the times
are deterministic, so that
. In our context, the random evolution
and its rescaled version
are simply equal to a functional of the spot price:
Because
, Theorem 8 tells us that the limit of
is trivial:
. In addition, we get:
so that by Theorem 11, we have the convergence in the Skorohod topology:
where
are orthogonal martingale measures on
with intensity
. This leads to the approximation:
Because the “first-order” limit is trivial, it calls to study the “second-order” limit, or diffusion limit, i.e., the convergence of
. This is indeed what is done in [
11]. A complete and rigorous treatment of the diffusion limit case for time-inhomogeneous random evolutions is below the scope of this thesis. This is done in the homogeneous case in [
10] (Section 4.2.2), or in a simplified context in [
6,
8]. Nevertheless, in the model we are presently focusing on, we can characterize the diffusion limit. The martingale representation of Lemma 5 is trivial as
. We define another rescaled random evolution
which is defined exactly as
, but with
replaced by
in the product. In our context:
Using the same techniques as in Lemma 5 (but going one more order in the “Taylor” expansion and writing
for suitable function
), we can get a martingale representation of the form:
for a suitable operator
. In fact, after a few lines of computation, we get:
In our present model,
is independent of time and the above reduces to:
With this notation, we have:
and therefore:
Denoting
for clarity, we get:
At this point, it can be noted that the above result should coincide with the quantity
of [
11] (Proposition 5.1), and it does! Note, however, that, because the times
are random in our context, the above quantity is multiplied by
in
, i.e., the inverse of a weighted average of the mean values of the inter-arrival times
. If
as in [
11], then
and so
: our framework is therefore a generalization of the one in [
11]. One interesting information coming from the above result is that in case of zero “fundamental correlations”
for
, we get for
, denoting
:
This is the point of the model: distressed selling (modeled by the function
g) induces a correlation between assets even if their fundamental correlation are equal to zero. It now remains to compute the quantities
. After some tedious computations, we get, denoting again
(and similarly for
):
The generator
is therefore given by:
We get in consequence, still denoting
for clarity:
and:
which is exactly equal to the result of [
11], Theorem 4.2, at the small exception that in their drift result for
, the
above denoted by
is replaced by a
, where we recall that
. We checked our computations and did not seem to find a mistake, therefore, to the best of our knowledge, the coefficient in
should indeed be
and not
. We will leave this small issue for further analysis. By Equation (
471), any limiting operator
of
satisfies:
Because of the above specific form of
, it can be proved as in [
11] (assuming
), using the results of [
20], that the martingale problem related to
is well-posed, and therefore using Equation (
492) that
converges weakly in the Skorohod topology as
to the
d-dimensional process
solution of:
where
W is a standard
d-dimensional Brownian motion and
. This result is a generalization of the one of [
11] as in our context, the times
at which the price jumps are random. The consequence of this randomness is that the driving coefficients
b and
c of the limiting diffusion process
are multiplied by, respectively,
and
. If
as in [
11], then
and we retrieve exactly their result. As mentioned above, our very general framework allows one to be creative with the choices of the operators
D and Γ, leading to possibly interesting extensions of the considered model. For example, one could simply “enlarge” the Markov chain
and consider the same model but with regime-switching parameters
and
, leading to diffusion limit results at almost no extra technical and computational cost.