Testing for jumps in near non-stationary diffusion processes



Sébastien Laurent1 and Shuping Shi2 1

Aix-Marseille University, CNRS & EHESS and Graduate School of Management 2

Macquarie University and CAMA November 17, 2017

Abstract In this paper, we show that despite the fact that Ornstein-Uhlenbeck (OU) processes fall within the general specification of asset price dynamics studied by Lee and Mykland (2008), the finite sample performance of their two tests for additive jumps is far from being satisfactory when the process deviates from the random walk, resulting in a strong size distortion and a dramatic power loss. Therefore, we propose a simple modification to their test that does not deteriorate its performance in the random walk case but improves the finite sample performance for local-to-unity processes (in both explosive or stationary directions). We apply the tests on 21 years of 5-minute logreturns of the Nasdaq stock price index and find that, unlike the other two tests, our test allows to detect jumps when log-prices exhibit clear upward or downward trend movements. Keywords: Jumps, Ornstein-Uhlenbeck processes, random walk, local-to-unity, intradaydata.

JEL classification: C12, C14. ∗ The authors gratefully acknowledge Jun Yu and participants at the econometric study group at the Singapore Management University for helpful discussions. Sébastien Laurent, Aix-Marseille School of Economics, Aix-Marseille University; Email: [email protected]. Shuping Shi, Department of Economics, Macquarie University and the Centre for Applied Macroeconomic Analysis; Email: [email protected].

1

1

Introduction

Testing for jumps and identifying precisely their occurrence is of overwhelming importance in finance because jumps have implications in risk management, portfolio allocation and derivatives pricing (A¨ıt-Sahalia, 2004). Several tests have been proposed in the literature (See Mancini and Calvori, 2012 for a survey). The most popular test is probably the one proposed independently by Andersen et al. (2007) and Lee and Mykland (2008, LM hereafter). The popularity of this test is due to its ease of implementation, its good size and power under some quite general conditions (for example, allowing both drift and volatility terms to be stochastic), the fact that it allows testing the presence of jumps within a short period of time (e.g., one day), and that it allows precise identification of jump arrival time. This test has been shown to have the overall best performance by Dumitru and Urga (2012) in a comprehensive Monte Carlo comparison among nine jump detection procedures available in the literature. Our paper provides important guidance on the implementation of the two tests proposed by Lee and Mykland (2008). We examine the performance of the LM tests under an Ornstein-Uhlenbeck (OU) diffusion process. The OU process is an important class of dynamics and has been studied extensively in the literature.1 The discretized OU process allows for random walk as well as local-to-unity dynamics (in both stationary and explosive directions) of log-prices in a high-frequency setting. The local-to-unity process, which was introduced by Phillips (1987), deviates mildly from the random walk. It bridges the asymptotic gap between the random walk and stationary and explosive processes. We demonstrate that the OU process satisfies the general requirements of asset price dynamics of Lee and Mykland (2008) and hence the limit theory of the LM tests is valid under this data generating process. Nevertheless, we find the finite sample performance of the LM tests to be unsatisfactory when the log-price process deviates from the random walk. Indeed, under this circumstance, this test is severely undersized which translates into a dramatic loss of power. To address this finite sample problem, we propose an 1 See, for example, Barndorff-Nielsen and Shephard (2001), Nicolato and Venardos (2003), Wang and Yu (2016), and Zhou and Yu (2015).

2

alternative construction of the test statistic. Despite its ease of implementation, our test improves the finite sample performance in the local-to-unity case. Furthermore, we demonstrate that the LM test is not applicable for mildly integrated processes (Phillips and Magdalinos, 2007). Compared to local-to-unity processes, mildly integrated processes deviate further away from the random walk and include the conventional stationary and explosive processes as special cases. To be valid, the LM tests require the drift and diffusion terms not to change dramatically over a short-time interval. These conditions are, unfortunately, not satisfied when the process is mildly integrated. Finally, as an illustration, we apply the three tests discussed in this paper on the 5minute log-returns of the Nasdaq stock index from 1996 to 2016. The main conclusion is that our test allows identifying more jumps. These additional jumps are embedded in periods with upward or downward trends when log-prices might have deviated from the random walk dynamics.

2

The Lee and Mykland Test

Let rt+jh = yt+jh −yt+(j−1)h be the j’s log-return of day t of an asset observed at the frequency h. By convention, we assume that we observe K = 1/h equidistant log-prices (yt+jh , j = {1, . . . , K}) per day. The J-statistic of Andersen et al. (2007) and Lee and Mykland (2008) is defined as Jt+jh = c1

rt+jh , sˆt

(1)

where c1 = E(|U |) = 0.7979 with U being the standard normal distribution and sˆt is an estimate of the instantaneous volatility based on bi-power variation2 such that

v u u sˆt = t

K

1 X |rt+jh ||rt+(j−1)h |. K −2 j=3

2

In the empirical application, we account for the presence of intraday periodicity in volatility when estimating the spot volatility, as in Andersen et al. (2007) and Boudt et al. (2011).

3

Lee and Mykland (2008) also propose a demeaned version of their statistic in the non-zero drift case, i.e., rt+jh − m ˆt J˜t+jh = c2 , sˆt where c2 = c1

p (K − 1) /(K − 2) and m ˆt =

1 K−1

PK

j=2 rt+jh

(2)

(i.e., the empirical mean of the

K − 1 intraday log-returns). Lee and Mykland (2008) show that under the null hypothesis of no jump, the test statistics (1) and (2) converge to a standard normal distribution as the sampling interval h goes to zero, provided that K is sufficiently large. Specifically, if K = Op (hα ) with −1 < α < −0.5, we have sup |Jt+jh − U | = Op (hη ) and sup |J˜t+jh − U | = Op (hη ), j

j

where − < η <

3 2

+ α −  with  ≥ 0. This limiting property holds for the following class of

diffusion processes: dyt = µ(t)dt + σ(t)dW (t),

(3)

where W (t) is an Ft -adapted standard Brownian motion with Ft : t ∈ [0, T ] being a rightcontinuous information filtration. The validity of the limit theory requires the following two conditions:

(1) sup j

(2) sup j

sup

µl − µt+(j−1)h = Op (h1/2− )

(4)

σl − σt+(j−1)h = Op (h1/2− )

(5)

t+(j−1)h≤l≤t+jh

sup t+(j−1)h≤l≤t+jh

for any  ≥ 0. In other words, the drift and diffusion coefficients should not change dramatically over a short time interval. This is satisfied for most Ito processes. It allows both the drift and stochastic volatility to depend on the log-price process itself. The jump test is implemented for each individual observations within the day. To control for the size of multiple tests, while Andersen et al. (2007) use a Bonferroni correction, 4

Lee and Mykland (2008) suggest using critical values based on the extreme value theory. Given that the maximum of a set of L i.i.d. realizations of the standard normal random variable is asymptotically a Gumbel distribution (see, for example, Aldous, 1989), i.e.,

maxj {Jt+jh } − CL → ξ, SL

(6)

where CL = (2 log L)1/2 − 21 (2 log L)−1/2 [log 4π +log(log L)], SL = (2 log L)−1/2 , and ξ is the standard Gumbel distribution with cumulative distribution function of P {ξ ≤ x} = exp [− exp(−x)]. The probability of maxj |Jt+jh | (over a set of L values) exceeding the critical value cvL,β is 100β% such that



 max |Jt+jh | > cvL,β

P

j

      cvL,β − CL =β = 2P max {Jt+jh } > cvL,β = 2 1 − exp − exp − j SL

and hence cvL,β = CL − SL log [− log (1 − β/2)] .

(7)

Controlling the overall size of the test each day of K observations can be done by setting L = K and therefore using the critical value cvK,β . The above test is also used to identify the location of jumps by declaring that there is a jump at time t + jh if |Jt+jh | > cvL,β . When L = K, the probability of finding at least one spurious jump (either positive and negative) within each day is therefore 100β%. There are two important corrections that we make for the implementation of this test. First, we believe there is a typesetting error in Lee and Mykland (2008) for the expression of CL , where the last part of CL should be [log 4π + log(log L)] instead of [log π + log(log L)]. Second, given that the jump test is a two-sided test, one should divide the confidence level by two when calculating the critical values. Specifically, we have cvL,β specified as in (7) instead of CL − SL log[log(1 − β)] as in Lee and Mykland (2008).3 3

The same applies to the demeaned LM test based on J˜t+jh .

5

2.1

The Ornstein-Uhlenbeck Diffusion Process

The OU diffusion process is

dyt = θ (yt − µ) dt + σt dWt

with constants θ and µ,

(8)

where σt is a Ft -predictable process satisfying (5) and Wt is the standard Wiener process. This process is a special case of (3) with µ(t) = θ (yt − µ). When θ = 0, the drift term µ(t) = 0 for all t and hence Condition (4) is satisfied. When θ 6= 0, Condition (4) requires that sup

sup

j

t+(j−1)h≤l≤t+jh

yl − yt+(j−1)h = Op (h1/2− )

(9)

for any  ≥ 0. To investigate the validity of this condition, we consider a discrete set of equally spaced data sampled at frequency h and {yt+jh : t = {0, 1, . . . , N − 1} , j = {1, . . . , K} with K = 1/h }. The total sample size T = N/h. Using Corollary 8.2.4 of Arnold (1974), we can show that the exact discrete solution of the OU process in (8) is

yt+jh = gh (θ) + αh (θ) yt+(j−1)h + Vt+jh (θ),

with gh (θ) = µ [1 − exp (θh)], αh (θ) = exp (θh), and Vt+h (θ) =

(10)

R t+jh

θ(t+jh−l) σ dW . l l t+(j−1)h e

The intercept gh (θ) converges to zero at a rate of h, i.e. gh (θ) = Op (h). When θ 6= 0, the autoregressive coefficient

αh (θ) = exp (θh) = 1 + θh + Op h2



converges to unity as h → 0 at a rate of h. The order of magnitude of the autoregressive coefficient Op (h) can be written as Op (1/T ) given that h = N/T and N is a constant. Therefore, the dynamic in (10) is the continuous time correspondence of the local-to-unity process of Phillips (1987), in the explosive direction when θ > 0 and stationary direction

6

when θ < 0.4 We refer to Model (8) with θ 6= 0 to as a near non-stationary5 diffusion process. When σ = (σs )s≥0 is locally bounded and satisfies Condition (5), we have Z

t+jh

e

Vt+jh (θ) =

θ(t+jh−l)



Z

t+jh

σl − σt+(j−1)h dWl +

eθ(t+jh−l) σt+(j−1)h dWl

t+(j−1)h

t+(j−1)h

Z

t+jh

eθ(t+jh−l) dWl {1 + op (1)}

= σt+(j−1)h t+(j−1)h

= σt+(j−1)h vt+jh {1 + op (1)} ,

where vt+jh is i.i.d normal with mean zero and variance (e2θh − 1)/(2θ). The variance of Vt+jh (θ) goes to zero as h → 0 and

lim

2 σt+(j−1)h (e2θh − 1)/(2θ)

h

h→0

2 = σt+(j−1)h .

Therefore, we have Vt+jh (θ) = Op (h1/2 ).

2.2

The Limiting Properties

Consider a limiting sequence {h = h1 , h2 , . . . , hn } such that Tn = N/hn is integer-valued with n o∞ Tn hn → 0 and Tn → ∞ as n → ∞ and a triangular array of random variables {yns }s=1 . n=1

When there is no confusion, we simply write hn as h, gh (θ) as gh , Vns (θ) as Vns , and αh (θ) as αh . The regression equation can be re-written as

yns = gh + αh yns−1 + Vns , 4

(11)

Notice that if we also allow N → ∞, the order magnitude of αh (θ), i.e. Op (N/T ), is smaller than the local to unity process (Op (1/T )) and hence the process is similar to the mildly integrated process of Phillips and Magdalinos (2007). 5 This term was also used in A¨ıt-Sahalia and Park (2012).

7

where s = t/h + j or equivalently

yns = gh

s−1 X

αhi

+

αhs yn0

i=0

+

s−1 X

αhi Vns−i .

(12)

i=0

Lemma 2.1 Under the data generating process (10) with θ 6= 0, yns = Op (1). The proof of Lemma 2.1 is in the appendix. Lemma 2.1 states that the asset price yns has an order of magnitude of Op (1) when the process deviates from the random walk dynamic in the form of (11). Given that gh = Op (h), αh − 1 = Op (h), and Vns = Op (h1/2 ), the log-return rns = yns − yns−1 has an order of magnitude of Op (h1/2 ) since rns = gh + (αh − 1) yns−1 + Vns

(13)

= Vns {1 + op (1)}

(14)

= Op (h1/2 ).

(15)

Consequently, Condition (9) is satisfied and hence the limiting theory of the LM tests remains valid for the OU process (8). Alternatively, (13) can be re-written as " rns = gh + (αh − 1) gh

s−2 X

# αhi

+

αhs−1 yn0

+ (αh − 1)

αhi Vns−1−i +

i=0

i=0

|

s−2 X

{z constant

}

|

{z AR

}

Vns . |{z} volatility

(16)

The constant component of the log-returns has an order of magnitude of Op (h) such that

" gh + (αh − 1) gh

s−2 X

# a

αhi + αhs−1 yn0 ∼ (αh − 1) exp (θt) (y0 − µ) .

i=0

When θ 6= 0, it converges to zero as h → 0. The second component, which arises from the autoregressive term (αh − 1) yns−1 in (13), is referred to as the AR component. From the proof of Lemma 2.1 in the appendix, we know that 8

Ps−2 i=0

αhi Vns−1−i follows a normal

distribution and hence the second term (αh − 1)

Ps−2 i=0

αhi Vns−1−i has an order magnitude

of Op (h). Since Vns = Op (h1/2 ), the log-return process is dominated by the volatility component asymptotically. In other words, the mean component (including both the constant and the AR terms) is asymptotically negligible. This is consistent with the argument of Barndorff-Nielsen and Shephard (2004) and Lee and Mykland (2008). Given the above results, one should expect identical performance of the J and J˜ (with demeaned return in the numerator) statistics when the data follows a random walk (θ = 0) or a local-to-unity process with ultra-high frequency (θ 6= 0 and h → 0). It is, however, important to note that in practice ultra-high frequency data of asset prices are often not available. Moreover, to avoid the impact of microstructure noise, it is a common practice to rely on lower frequency data, such as 5 or 10 minutes when applying the LM test (Park and Linton, 2011). Due to these practical constrains, the mean of relatively low frequency (e.g., 5-minute) log-returns could be of non-negligible magnitude when the process deviates from the random walk. Therefore, we might expect the LM test to have size distortion in finite samples when θ 6= 0.

2.3

The Finite Sample Performance of the Lee and Mykland Tests

In this section, we investigate the finite sample performance of the LM test statistics (1) and (2). The size of the tests is examined under the data generating process (10). Under the alternative hypothesis, we extend (10) by allowing exactly one additive jump per day such that yt+jh = gh (θ) + αh (θ) yt+(j−1)h + φt+h It+jh + Vt+jh (θ),

(17)

where It+jh is a dummy variable indicating the jump location (whose occurrence is random) and φt+h the jump size. The error term Vt+h (θ) is specified as √ Vt+h (θ) = σt+jh hεt+jh 2 2 σt+jh = α0 + σt+(j−1)h (β1 + α1 vt+jh ),

9

(18) (19)

i.i.d.

where εt+jh and vt+jh are two independent ∼ N (0, 1) random variables. Equation (19) is a standard Euler discretization of the GARCH(1,1) diffusion process of Nelson (1991), i.e.,

dσt2 = κ[ω − σt2 ]dt +



2λκσt2 dwt ,

(20)

where κ > 0, ω > 0, and 0 < λ < 1 are parameters and wt is a Wiener process. Parameters √ of (19) are linked to (20) as follows: α0 = κωh, β1 = 1 − κh, α1 = 2λκh. We consider a wide range of values for θ, nesting unit root as well as local-to-unity in both the explosive and stationary directions. The remaining parameters are set as in Laurent and Shi (2017), i.e., y0 = 6.959, µ = 0.0002. For the volatility, we follow Andersen and Bollerslev (1998) and choose the parameters κ = 0.035 and λ = 0.296 to simulate a realistic log-price process with very persistent GARCH effects and set ω = 0.012 such that 2 ) = 0.012 . Under the alternative, we consider a negative jump6 of magnitude 60% E(σt+jh

times the spot volatility, i.e., φt+h = −0.6σt+h . In all simulations, we generate 24, 000 observations, corresponding to one day (N = 1) of 1-second data of a stock or index traded about 6.5 hours.7 This implies that T = K and h = 1/K. We then aggregate the simulated log-prices at a lower frequency by picking-up one observation every 60 or 300 data points to get respectively 1-minute (K = 400) and 5-minute (K = 80) data. Figure 1 plots one typical sample path of 5-minute log-prices of the OU process without jumps and with N = 1 and θ ranging between −0.05 to 0.05.8 This figure illustrates the fact that when θ deviates from 0, simulated log-prices depict clear upward or downward trends. The empirical performance of the LM statistics (1) and (2), denoted respectively by J ˜ is reported in Table 1 for three sampling frequencies, i.e., 1-second (K = 24, 000), and J, 1-minute (K = 400) and 5-minute (K = 80) and 106 replications. For the critical values, we set L = K and β = 0.1% so that the probability of finding at least one spurious jump within each day is 0.1% or equivalently, we expect one out of 1, 000 days containing at least 6

Simulation results are qualitatively the same for positive jumps and jumps of a random sign. This is the case of the NASDAQ stock market, trading from 9:30 to 16:00. 8 We do not consider θ values outside this range due to the local-to-unity property of the OU process. 7

10

Figure 1: A typical sample path of one day of 5-minute log-prices of the OU process with GARCH effects, no jump and with θ ranging between −0.05 and 0 (left panel) and 0 and 0.05 (right panel). θ=0.05 θ=0

6.95

7.30 θ=−0.01

6.90

θ=0.04

7.25 θ=−0.02

6.85

7.20 θ=0.03

7.15

6.80

θ=−0.03

6.75

7.10

θ=0.02

7.05

θ=0.01

θ=−0.04

6.70 θ=−0.05

7.00

6.65

0

20

40

60

80

6.95 0

θ=0

20

40

60

80

one spurious jump. The critical values used in this simulation are therefore respectively cv24000,0.1% = 5.644, cv400,0.1% = 5.033 and cv80,0.1% = 4.851. Notice that the asymptotic critical values (7) depend on the values chosen for both L and β. Here, we control for the overall size of the test over a day (L = K) and set β = 0.1%. Alternatively, we could control the size of the test over a longer period, eg. a year. Interestingly, when setting L = 250K and β = 5%, the critical value is 4.735 for 5-minute data and is therefore very similar to the one of the above setting (i.e., 4.851). We report the number of times that at least one jump is detected within the sample period. As highlighted (in bold) in the table, when the asset price follows a random walk dynamics (i.e., θ = 0), the original LM tests J and J˜ have an outstanding performance, with an empirical size slightly above 0.1%. Similar results have been obtained i) when volatility is constant (e.g. between 0.02% and 0.05% for both tests) and ii) for higher critical levels (i.e., between 4.1 and 4.9 when β = 5% for both tests) but not reported to save place. The power is 100% for K = 400 and K = 24, 000 and about 57% for K = 80. As expected, in absence of microstructure noise, jumps are easier to detect when the sampling frequency gets higher. However, the assumption of no microstructure noise for data sampled at the

11

1-second and even 1-minute frequency is unrealistic. Therefore, the LM tests are usually applied to 5-minute data to reduce the impact of microstructure noise, at the cost of power loss. Table 1: Empirical performance of the Lee and Mykland Test for a nominal size of .01 θ Size K = 24, 000 J J˜ K = 400 J J˜ K = 80 J J˜

-0.05

-0.04

-0.03

-0.02

-0.01

0.0

0.01

0.02

0.03

0.04

0.05

0.048 0.026

0.054 0.037

0.060 0.046

0.053 0.048

0.060 0.059

0.062 0.061

0.058 0.055

0.055 0.049

0.061 0.046

0.056 0.036

0.047 0.022

0.000 0.000

0.000 0.000

0.000 0.000

0.001 0.000

0.021 0.006

0.034 0.033

0.019 0.006

0.001 0.000

0.000 0.000

0.000 0.000

0.000 0.000

0.000 0.000

0.000 0.000

0.000 0.000

0.000 0.000

0.002 0.000

0.034 0.029

0.002 0.000

0.000 0.000

0.000 0.000

0.000 0.000

0.000 0.000

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

100.0 100.0

99.8 100.0

81.2 98.5

12.3 62.5

0.0 0.0

0.4 0.0

35.0 13.5

58.2 55.8

4.2 12.7

0.0 0.0

0.0 0.0

0.0 0.0

0.0 0.0

Power: with one jump per day K = 24, 000 J 100.0 100.0 J˜ 100.0 100.0 K = 400 J 98.8 100.0 J˜ 76.4 99.1 K = 80 J 0.0 0.0 J˜ 0.0 0.0

Importantly, although the LM tests are asymptotically valid under (10), both tests are strongly undersized in finite samples when the process deviates from the random walk, especially when the sampling frequency is low (i.e. K = 400 and 80). Indeed, both test almost never reject the null (while the null of no jump is true) when |θ| ≥ 0.02.9 This translates into a dramatic loss of power for both tests when K = 80 and |θ| ≥ 0.02, values for which both tests have a power close to 0. This verifies our conjecture that, when the log-price process deviates from the random walk, the non-negligible mean of log-returns might affect the performance of the LM tests in finite samples. Surprisingly, from Table ˜ does not improve the performance of the test. 1, the demeaned version of the test (i.e., J) 9

A close-to-zero rate of rejection is also observed when using a higher critical value of β = 5%.

12

3

A Modified Lee and Mykland Test

There are two potential reasons that the demeaned test proposed by LM does not solve the finite sample issue. First, the empirical mean of the K − 1 intraday log-returns (i.e., m ˆ t ) is not an accurate estimator of the drift component due to the influence of jumps. For the same reason that bi-power variation is used to estimate the spot volatility instead of the realized variance (in the denominator of the LM statistics), a jump-robust estimator of the mean is required for the numerator of the J˜ statistics. We propose using the median, denoted by m ˆ ∗t , instead of the empirical mean for the construction of the test. While the empirical mean has a breakdown point of 0% (a single large observation can throw it off), the median has a breakdown point of 50% and is therefore robust to jumps. Unreported simulation results suggest that this correction slightly improves the power of the test but, as expected, has no impact on its empirical size. Second, although the J˜ statistic corrects for the non-zero mean of log-returns in the numerator, it does not consider the impact of non-zero mean of log-returns on the volatility estimator in finite samples. Therefore, we propose a new volatility estimator based on centered log-returns in the next subsection.

3.1

A New Estimator for the Spot Volatility

To improve the finite sample accuracy of the volatility estimator, we propose using centered ∗ ∗ log-returns rt+jh for the calculation of the bi-power variation. Let rt+jh = rt+jh − m ˆ ∗t with

m ˆ ∗t = median(rt+2h , . . . , rt+Kh ) being the median of the K − 1 intraday log-returns. The new daily volatility estimator is denoted by sˆ∗t and defined as v u u ∗ sˆt = t

K

1 X ∗ ∗ |rt+jh ||rt+(j−1)h |. K −2 j=3

As illustrated in Section 2.2, both the constant and AR components in (16) are asymptotically dominated by the volatility component and hence using centered log-returns

13

∗ rt+jh for the calculation of bi-power volatility will not alter its limiting property. Given

(14), the asymptotic spot volatility of the return series is identical to that of Vt+jh , i.e. R t+jh

2θ(t+jh−l) σ 2 dl. l t+(j−1)h e

The integrated variance of day t (denoted IVt ) is therefore

Z

t

e2θ(t−l) σl2 dl.

IVt ≡ t−1

Table 2 reports the biases and root mean square errors (RMSE) of



K − 2ˆ st /c1 and



K − 2ˆ s∗t /c1 for the three sampling frequencies, measuring distances between the volatil√ ity estimators and the asymptotic volatility IVt . The data generating process is as in

Section 2.3 but without jumps. First, the biases and RMSEs of both estimators become smaller as the sampling frequency increases (i.e., when K gets larger). This is because while the return volatility arises from both the AR and Vns components (16), the volatility sourced from the AR term diminishes and is dominated by that of Vns as h → 0. Table 2: Estimation accuracy (bias and RMSE) of volatility estimators θ

-0.05

-0.04

-0.03

-0.02

-0.01

0.0

0.01

0.02

0.03

0.04

0.05

0.6 0.8

-0.0 0.5

-0.4 0.7

-0.6 0.8

-0.6 0.8

-0.4 0.7

0.1 0.6

6.1 7.4

-0.3 4.1

5.2 6.8

22.4 23.0

49.6 50.0

84.4 84.7

124.4 124.6

26.9 29.0

-1.7 9.2

26.8 29.0

99.6 100.6

192.5 193.1

291.2 291.6

391.4 391.7

0.5 0.7

-0.0 0.5

-0.5 0.7

-1.0 1.1

-1.5 1.6

-2.0 2.1

-2.6 2.6

-0.1 4.0

-0.6 4.1

-1.1 4.2

-1.6 4.3

-2.0 4.5

-2.4 4.7

-2.9 5.0

-2.1 9.1

-3.2 9.5

-3.1 9.5

-3.1 9.5

-2.9 9.4

-2.8 9.4

-2.6 9.5



st /c1 Bi-power variation estimator K − 2ˆ K = 24, 000 Bias×104 4.8 3.5 2.4 1.4 RMSE ×104 4.9 3.5 2.4 1.5 K = 400 Bias×104 119.2 82.4 49.7 23.4 RMSE ×104 119.3 82.6 50.0 23.9 K = 80 Bias×104 371.5 279.1 186.4 97.7 RMSE ×104 371.8 279.5 186.9 98.6 √ Bi-power variation estimator K − 2ˆ s∗t /c1 K = 24, 000 Bias×104 2.4 2.0 1.5 1.0 RMSE ×104 2.5 2.0 1.6 1.1 K = 400 Bias×104 2.1 1.6 1.0 0.4 RMSE ×104 4.6 4.3 4.2 4.1 K = 80 Bias×104 2.3 1.2 -0.0 -1.1 RMSE ×104 9.4 9.0 8.9 9.0

Second, when θ = 0, there is no obvious difference in the estimation accuracy (measured by bias and RMSE) between the two volatility estimators. The sˆ∗t estimator is, how-

14

ever, expected to improve the estimation accuracy of volatility in finite sample when θ deviates from zero. As we can see, it provides much more accurate estimation of the volatility than sˆt when |θ| = 6 0. While the bias and RMSE of sˆt increase consistently as θ deviates further away from zero, the estimation accuracy of sˆ∗t remains roughly unchanged. For example, when K = 400, the bias of sˆt rises from −0.3 × 10−4 to 1.119 × 10−2 when θ changes from 0 to −0.05. In contrast, the bias of sˆ∗t stays roughly the same (with an order of magnitude of 10−5 ).

3.2

A Modified Lee and Mykland Test

Based on findings above, we propose a new jump test statistic such that

∗ = c2 Jt+jh

rt+jh − m ˆ ∗t . ∗ sˆt

(21)

The finite sample performance of this test, denoted J ∗ , is illustrated in Table 3. The data generating processes and parameter settings are the same as in Section 2.3. Table 3: Empirical performance of the J ∗ statistic θ K = 24, 000 Size Power K = 400 Size Power K = 80 Size Power

-0.05

-0.04

-0.03

-0.02

-0.01

0.0

0.01

0.02

0.03

0.04

0.05

0.036 100.0

0.040 100.0

0.042 100.0

0.038 100.0

0.037 100.0

0.042 100.0

0.040 100.0

0.034 100.0

0.043 100.0

0.042 100.0

0.039 100.0

0.030 100.0

0.034 100.0

0.032 100.0

0.033 100.0

0.034 100.0

0.037 100.0

0.033 100.0

0.031 100.0

0.034 100.0

0.035 100.0

0.034 100.0

1.110 56.0

0.033 57.0

0.035 58.2

0.039 59.1

0.046 60.2

0.051 60.5

0.045 59.0

0.034 58.0

0.034 56.9

0.043 55.7

3.261 54.6

As we can see in Table 3, when the sampling frequency is high (i.e. K ≥ 400), the J ∗ test has an outstanding performance, i.e., for all values of θ, the empirical size is very close to the nominal level of 0.1% and the empirical power is 100%. When the sampling frequency is lower (i.e., five minutes or K = 80), the J ∗ test has very good size and power when |θ| ≤ 0.04. The power of the test (that is not adjusted for the size distortion) varies between 54% and 60% and is not very sensitive to the value of θ. Recall from Table 1 that for θ = 0 and K = 80, the empirical power of the J and J ∗ tests are found to be about 15

57% and that both tests suffer from serious size distortion towards 0 and lack of power for K = 80 and |θ| = 6 0. Interestingly, when K = 80, there is an obvious discontinuity in the empirical size of the test between |θ| = 0.04 and |θ| = 0.05 (italic). Indeed, the size of the test increases from about 0.04% for |θ| = 0.04 to 1.11% for θ = −0.05 and 3.26% for θ = 0.05. This size distortion, however, disappears when the sampling frequency increases from five minutes to one minute and above. This result is discussed in detail in the next subsection.

3.3

Discussion

Recall that the LM tests require the drift and diffusion coefficients in (3) not to change dramatically over a short time interval. We have shown above that the OU diffusion process (8), which is equivalent to a local-to-unity process, satisfies the required conditions, i.e. (4) and (5). Here, we show that a mildly integrated process of Phillips and Magdalinos (2007) does not satisfy these requirements. It is sufficient to assume that the error term has a constant volatility. Consider the following data generating process

i.i.d

yns = δh yns−1 + ωh εns , with εns ∼ N (0, 1) ,

(22)

√ where δh = 1 + θhψ with ψ ∈ [0, 1) and ωh = σ h.10 The autoregressive coefficient can be rewritten as δh = 1 + ζTn−ψ with ζ = θN ψ . Notice that the value of ψ is bounded above by one but does not include one (the case of local-to-unity). When ψ = 0, it becomes a stationary process with ζ < 0 or explosive process with ζ > 0. One can deduce that

yns = δhs yn0 + ωh

s−1 X j=0

10

From simplicity, we do not include a drift term.

16

δhj εns−j .

(23)

Lemma 3.1 Under the data generating process of (22), we have yns = Op (exp(hψ−1 )). The proof of Lemma 3.1 is in the appendix. Lemma 3.1 states that yns has an order of magnitude of Op (exp(hψ−1 )). Unlike the OU process, log-returns of a mildly integrated process have an order of magnitude of Op (hψ exp(hψ−1 )), namely rns = (δh − 1) yns−1 + ωh εns

(24)

= (δh − 1) yns−1 {1 + op (1)}

(25)

= Op (hψ exp(hψ−1 )).

(26)

The second equality comes from the fact that δh − 1 = Op (hψ ), ωh εns = Op (h1/2 ), and the result in Lemma 3.1. Returns in (24) are dominated by the drift component (δh − 1) yns−1 . This is compatible with the type of processes considered by Christensen et al. (2016), where the drift term has an order magnitude of Op (hζ ) with 0 < ζ < 1/2 and prevails over volatility. They are, however, different on the volatility side. While the volatility of the mildly explosive process (22) is assumed to be Op (h1/2 ), Christensen et al. (2016) allows for exploding volatility with an order magnitude of Op (h1/2−ζ ). Importantly, since the order of magnitudes of rns of both processes are larger than Op (h1/2 ), Condition (4) of the LM tests is violated. As such, all the LM-type of tests discussed in this paper (including J ∗ ) are not appropriate for either the mildly integrated processes or the process of Christensen et al. (2016). Furthermore, compared with the local-to-unity process (10), the mildly integrated process deviates further away from the random walk. With θ = −0.05 and K = 80, the coefficient αh = exp(θh) takes a value of 0.999375 which is much smaller than 0.999998 when K = 24, 000. Therefore, it is likely to be observationally equivalent to data generated from an mildly explosive process for which the conditions needed to apply the LM test do not apply. This serves as an explanation for the discontinuity in the performance of the LM test J ∗ in Table 3.

17

4

Empirical Application

In this section, we apply the three tests for jumps presented above on 5-minute data of the Nasdaq stock price index on the period spanning from 1996 to 2016. All trades before 9:30 am or after 4:00 pm are discarded as well as the first trade after 9:30 am, which is the usual way of avoiding the overnight effect. The choice of this series is dictated by the fact that several studies (see Phillips et al., 2011, Homm and Breitung, 2012, and Shi and Song, 2015, among others) have shown evidence of explosiveness in weekly and monthly data of the Nasdaq in the late 1990’s. The logarithmic prices are plotted year-by-year in Figure 2. For ease of exposition, we ignored so far intraday periodicity effects in the spot volatility. However, it is well known since the works of Taylor and Xu (1997) and Andersen and Bollerslev (1998b) that opening, lunch and closing of financial markets induce a strong periodic pattern in the volatility of high-frequency returns. More recently, Boudt et al. (2011) have proposed both a parametric and a nonparametric estimator of the intraday (or intraweek) periodicity that is robust to jumps as well as a correction to the LM jump statistics where the spot volatility depends on the estimated periodicity. They show that this modification helps to increase the power to detect the relatively small jumps occurring at times for which volatility is periodically low and reduces the number of spurious jump detections at times of periodically high volatility. Therefore, in the empirical application, we follow Boudt et al. (2011) in multiplying sˆt in (1) and (2) and sˆ∗t in (21) by the weighted standard deviation (WSD) estimate of the intraday periodicity. We refer to Boudt et al. (2011) for a detailed presentation of this estimator. Note that, to be consistent with the findings of Section 3.1, unlike in Boudt et al. (2011), we ∗ compute WSD with centered 5-minute log-returns (i.e., rt+jh ). The periodicity is estimated

year-by-year with a cycle length of one week (to allow different day-of-the-week effects) as in Boudt et al. (2011) and Lahaye et al. (2011) and standardized such that the squared periodicity factors average to one every day.11 11

A graph of the estimated periodicity is available upon request but not reported here to save space.

18

Figure 2: 5-minute log-prices of the Nasdaq stock index over the period 1996-2016 7.2

1996

7.1

7.4

7.0

7.2

6.9 0

4800

9600

14400

7.6

1998

7.4 4800

9600

14400

8.5

1999

8.25

0

1997

0 8.00

4800

9600

14400

9400

14100

9800

14700 19600

9800

14700

7.75

8.00

8.0

7.75 0 7.75

4900

9800

14700 19600

2002

0

7.6

7.50 2000

4900

9800

14700

2003

7.50

9600

14400 19200

2005

7.7

0

7.8

4700 2004

7.5

7.2 4800

2001

7.6

7.4 7.25 0

7.25 0 7.7

4850

9700

14550

2006

0 8.0

4900 2007

7.9 7.6

7.7 7.8

0

8.00

5000

10000 15000

0

7.75

2008

7.75

4950

9900

14850

7.9

2009

7.50

0

9900

14850

0

2011

14850

5000

10000 15000

0

4950

9900

14850

10000 15000

2013

8.1 4850

9700

14550

0 8.6

8.5

8.5

8.4

8.4

0

5000

8.2

2015

8.4

0 8.3

7.9

2014

8.3 0

9900

8.0

7.8 0

4950 2012

7.9

2010

7.7

7.25 4950

4900

7.8

7.50

7.25

0

4950

9900

19

14850

0

4900

9800

14700 19600

9300

13950

2016

4650

Empirical results concerning the tests for jumps are reported in Table 4. Columns J, J˜ and J ∗ correspond respectively to the jump statistics (1), (2) and (21). For the critical value cvL,β , like for the simulations, we set L to the total number of 5-minute log-returns per day, i.e., L = 78 and β = .1%. The values reported in the three columns below ‘Significant jumps’ correspond to the number of jump statistics greater than the critical value while those below ‘Significant days’ correspond to the number of days for which at least one significant jump is detected. The first conclusion we can draw from Table 4 is that for each year, the numbers of jumps detected by the three tests are not dramatically different. However, our proposed J ∗ statistic allows to detect almost systematically more jumps. Over the 21 years, 439, 417 and 475 jumps are detected using respectively the statistics (1), (2) and (21), which corresponds respectively to 402, 385 and 436 days with at least one significant jump. This is consistent with our simulation findings that, when the log-price process slightly deviates from the random walk, the newly proposed test J ∗ is less undersized and has higher power than the other two tests. Note also that much fewer jumps are detected using the J˜ statistic. We attribute this difference to the non robustness to jumps of the estimator of the mean used in the numerator. This leads to a failure to detect small jumps, a phenomenon called ‘outlier masking’ in the robustness literature (see Davies and Gather, 1993).12 To better understand the difference between the tests, the log-prices of the 38 days for which a jump is detected only by the J ∗ statistic, i.e., J ∗ > cv78,0.1% while J and J˜ ≤ cv78,0.1% are plotted in Figures 3 and 4. Interestingly, these days are characterized by a strong upward or downward trend compatible with those plotted in Figure 1 that we attribute to a temporary deviation of the random walk hypothesis. Recall from Tables 1 and 3, that when the dynamics of log-prices deviates marginally ˜ applied on 5-minute data are from the random walk, the original LM tests (both J and J) found to be undersized and have no power while our test has good size and very good power. However, when the dynamics of log-prices deviates dramatically from the random When the median is used instead of the mean in the J˜ statistic, the number of detected jumps is approximatively the same as the one of the J statistic. Results are not reported to save space but available upon request. 12

20

Figure 3: Log-prices of days for which a jump identified by J ∗ but not J and J˜ 1997-01-27

1997-02-19

7.22

7.2225

7.21

7.2175 0

50

0

2000-03-15

1997-10-06

0

8.285

8.20 0

7.96 7.94 0

7.5525

0

50

0

2005-02-08

50

50

0

2005-10-10

7.760

7.640 50

7.758 0

50

0

2007-09-05

7.82

7.870

7.81

7.865

50 2006-05-08

7.645

0

50 2004-05-21

7.14

2007-02-09

7.765

0 7.5575

50

7.770

50 2004-03-30

7.595

2006-10-31

0

50

50 2000-11-22

7.600

7.6475 7.6450 7.6425

7.555

0

7.16

50 2004-10-26

50 2000-09-27

2002-10-18

7.49

0

0 8.22

50

0

7.33

8.295

2002-04-18

7.565

7.450 50

8.44

7.51

7.34

2000-08-22

8.46

1997-12-15

7.452

50 2008-02-01

7.785 7.775

0

50

0

50

0

50

Note: The vertical line indicates the arrival time of detected jumps.

21

˜ con’t Figure 4: Log-prices of days for which a jump identified by J ∗ but not J and J, 2008-03-31

2010-02-11

7.69

7.735

2011-05-09

7.67

7.725 0

50

0

2011-06-29

7.9175

8.030

7.9125

8.025 0

50

7.885

7.950

7.880

50 2012-10-11

0

0

50 8.4325

8.190

8.2725

8.4300

50 2014-10-06

8.41

50

0

2015-06-08

50 2015-06-23

8.530

0

50 2015-10-12

8.486

8.548

8.525 50

0

8.40 0

2014-10-30

0

50 2014-09-08

8.2775

50 2013-06-07

8.145 0

8.195

8.430 8.425 8.420

0 8.150

2013-11-01

50

50 2013-03-13

8.0875 8.0825

2013-08-21

0

2011-06-13

7.955

8.482

8.546 0

2016-08-12

50

0

50

0

50

2016-08-31

8.560

8.562

8.556

8.560 0

50

0

50

Note: The vertical line indicates the arrival time of detected jumps.

22

Table 4: Descriptive statistics on significant jumps Significant jumps

1996 1997 1998 1999 2000 2001 2002 2003 2004 2005 2006 2007 2008 2009 2010 2011 2012 2013 2014 2015 2016 All

J 18 17 23 14 12 18 20 23 20 39 31 25 14 27 24 18 25 18 17 18 18 439

J˜ 18 15 21 14 11 18 21 23 19 37 26 25 12 22 23 19 21 20 17 18 17 417

J∗ 19 19 24 14 17 18 24 22 22 41 32 27 14 24 27 22 25 24 20 21 19 475

Significant days J 16 15 20 12 11 16 18 21 19 35 27 23 13 25 24 17 24 17 16 16 17 402

J˜ 16 13 19 12 11 16 19 21 18 34 23 23 11 21 23 18 20 19 16 16 16 385

J∗ 17 17 22 12 16 16 21 21 21 37 28 24 12 23 26 21 24 22 19 19 18 436

walk, all three tests for jumps considered in this paper do not perform well on 5-minute data. For example, when K = 80 and |θ| = 0.05, the original LM tests are undersized (with an empirical size of zero) and have no power while our J ∗ test is slightly oversized. A natural question arises on whether all uniquely identified jumps via J ∗ are true jumps or false jumps due to a strong deviation from the random walk. Note first of all that we would expect to find more jumps with J ∗ than with the other two tests in the presence of large deviations of the random walk. The period 1996 to 1999 is the one for which the literature reports the most evidence of explosiveness in the Nasdaq stock price (see Phillips et al., 2011 among others). However, during this period, we only find four additional jumps over the four-year interval suggesting that if the process has indeed episodes of

23

explosiveness during this period, it falls in the category of local-to-unity processes.13 A second argument suggesting that the jumps uniquely identified via J ∗ are truly jumps is the visual inspection of the log-prices around the interval corresponding to the significant jumps (highlighted by a vertical line in Figures 3 and 4). These figures clearly suggest that log-prices are characterized by a large discontinuity around the detected jumps that is not detected by the other two tests due to the presence of a strong upward or downward trend.

5

Conclusion

In this paper, we examine the performance of the LM tests under an Ornstein-Uhlenbeck diffusion process, whose exact discrete time solution nests random walk as well as localto-unity (explosive or stationary) processes. We show that despite the fact that OU processes fall within the general specification of asset price dynamics studied by Lee and Mykland (2008), the finite sample performance of the LM tests is far from being satisfactory when the process deviates from the random walk, resulting in a strong size distortion and a power loss. To address this finite sample problem, we propose an alternative construction of the test statistic, which significantly improves its performance in the local-to-unity case and does not affect its performance in the random walk case. The newly proposed test as well as the two tests proposed by Lee and Mykland (2008) are applied on 5-minute logreturns of the Nasdaq on the period spanning from 1996 to 2016. Slightly more jumps are detected using our test statistic. Interestingly, on days where jumps are detected only with our test, log-prices exhibit clear upward or downward trend movements. We attribute these trends to a temporary deviation of the random walk hypothesis explaining why the original tests of Lee and Mykland (2008) fail to detect jumps on these days.

13

Christensen et al. (2016) develop a test for identifying episodes of drift bursting which is similar to testing the presence of large deviations in the asset price dynamics.

24

Appendix Proof. [Proof of Lemma 2.1] The first two terms of (12)

gh

s−1 X

a

αhi + αhs yn0 ∼ µ + exp (θt) (x0 − µ) ,

(27)

i=0

where s = t/h + j. The variance of the third term s−1 X

V ar

s−1 X

! αhi Vns−i

= V ar

i=0

! αhi σns−i−1 vns−i {1 + op (1)}

i=0 s−1

=

e2θh − 1 X 2i 2 αh σns−1−i = Op (h)Op (h−1 ) = Op (1) 2θ i=0

when (σt )t≥0 is locally bounded. Therefore, by the Central Limit Theorem for weighted sums, the third term of (12) converges to a normal distribution. Therefore, yns = Op (1). Proof. [Proof of Lemma 3.1] The first term of (23) δhs yn0

ψ s

ψ

= (1 + θh ) y0 = exp(sθh )y0 = exp



  t ψ + j θh y0 h

  = exp tθhψ−1 + jθhψ y0 = exp(hψ−1 ) exp(tθ + jθh)y0

which diverges to infinity at a rate of Op (exp(hψ−1 )) as h → 0. For the second term of (23), we have  E ωh

s−1 X

2 δhj εns−j  = ωh2

j=0

s−1 X

δh2j

j=0

= ωh2

δh2s − 1 δh2 − 1 25

=

exp ωh2

 2tθhψ−1 + 2jθhψ − 1 2θhψ + θ2 h2ψ

  ω 2 exp(tθ + 2jθh) − 1 = exp 2hψ−1 h1−ψ h h 2θ + θ2 hψ   exp(tθ) − 1 a ∼ exp 2hψ−1 h1−ψ σ 2 . 2θ By CLT for weighted sums, we have



ψ−1

exp h



h

1−ψ 2

ωh

s−1 X

δhj εns−j

j=0

  2 exp(tθ) − 1 . → N 0, σ 2θ L

Therefore, the first term dominates the second term. The process yns has an order of magnitude of Op (exp(hψ−1 )).

26

References A¨ıt-Sahalia, Y., 2004. Disentangling diffusion from jumps. Journal of Financial Economics 74, 487–528. A¨ıt-Sahalia, Y., Park, J. Y., 2012. Stationarity-based specification tests for diffusions when the process is nonstationary. Journal of Econometrics 169 (2), 279–292. Aldous, D., 1989. Probability approximations via the poisson clumping heuristic. Vol. 77. Springer-Verlag, New York. Andersen, T., Bollerslev, T., 1998. Answering the skeptics: Yes, standard volatility models do provide accurate forecasts. International economic review, 885–905. Andersen, T. G., Bollerslev, T., 1998b. Deutsch mark-dollar volatility: intraday activity patterns, macroeconomic announcements, and longer run dependencies. Journal of Finance 53, 219–265. Andersen, T. G., Bollerslev, T., Dobrev, D., 2007. No-arbitrage semi-martingale restrictions for continuous-time volatility models subject to leverage effects, jumps and iid noise: Theory and testable distributional implications. Journal of Econometrics 138 (1), 125–180. Arnold, L., 1974. Stochastic differential equations. A Wiley-Interscience publication. Wiley. Barndorff-Nielsen, O. E., Shephard, N., 2001. Non-gaussian ornstein–uhlenbeck-based models and some of their uses in financial economics. Journal of the Royal Statistical Society: Series B (Statistical Methodology) 63 (2), 167–241. Boudt, K., Croux, C., Laurent, S., 2011. Robust estimation of intraweek periodicity in volatility and jump detection. Journal of Empirical Finance 18 (2), 353–367. ` R., 2016. The drift burst hypothesis. Working paper, Christensen, K., Oomen, R., Reno, Aarhus University.

27

Davies, L., Gather, U., 1993. The identifcation of multiple outliers. Journal of the American Statistical Association 88, 782–792. Dumitru, A., Urga, G., 2012. Identifying jumps in financial assests: a comparison between nonparametric jump tests. Journal of Business and Economic Statistics 30, 242–255. Homm, U., Breitung, J., 2012. Testing for speculative bubbles in stock markets: a comparison of alternative methods. Journal of Financial Econometrics 10 (1), 198–231. Lahaye, J., Laurent, S., Neely, C., 2011. Jumps, cojumps and macro announcements. Journal of Applied Econometrics 105, 893–921. Laurent, S., Shi, S., 2017. Testing for jumps in near non-stationary diffusion processes. Tech. rep., Available at SSRN: https://ssrn.com/abstract=3052053. Lee, S., Mykland, P., 2008. Jumps in financial markets: A new nonparametric test and jump dynamics. Review of Financial Studies 21 (6), 2535–2563. Mancini, C., Calvori, F., 2012. Jumps. In: Bauwens, L., Hafner, C., Laurent, S. (Eds.), Handbook of Volatility Models and Their Applications. Wiley. Nelson, D. B., 1991. ARCH models as diffusion approximations. Journal of Econometrics 45, 7–38. Nicolato, E., Venardos, E., 2003. Option pricing in stochastic volatility models of the ornstein-uhlenbeck type. Mathematical Finance 13 (4), 445–466. Park, S., Linton, O., 2011. Realized volatility: Theory and application. In: Bauwens, L., Hafner, C., Laurent, S. (Eds.), Handbook of Volatility Models and Their Applications. Wiley. Phillips, P. C., 1987. Time series regression with a unit root. Econometrica, 277–301. Phillips, P. C., Magdalinos, T., 2007. Limit theory for moderate deviations from a unit root. Journal of Econometrics 136 (1), 115–130.

28

Phillips, P. C., Wu, Y., Yu, J., 2011. Explosive behavior in the 1990s NASDAQ: When did exuberance escalate asset values? International Economic Review 52 (1), 201–226. Shi, S., Song, Y., 2015. Identifying speculative bubbles using an infinite hidden Markov model. Journal of Financial Econometrics 14 (1), 159–184. Taylor, S., Xu, X., 1997. The incremental volatility information in one million foreign exchange quotations. Journal of Empirical Finance 4, 317–340. Wang, X., Yu, J., 2016. Double asymptotics for explosive continuous time models. Journal of Econometrics 193, 35–53. Zhou, Q., Yu, J., 2015. Asymptotic theory for linear diffusions under alternative sampling schemes. Economics Letters 128, 1–5.

29

Testing for jumps in near non-stationary diffusion ...

Nov 17, 2017 - statistics (1) and (2) converge to a standard normal distribution as the sampling ...... Journal of Business and Economic Statistics 30, 242–255.

1MB Sizes 2 Downloads 164 Views

Recommend Documents

Hellinger Distance Based Drift Detection for Nonstationary ... - CiteSeerX
Detecting change in data distributions is an important prob- lem in data mining and machine learning algorithms due to increasing number of applications that ...

Symmetry Breaking by Nonstationary Optimisation
easiest to find under the variable/value order- ing but dynamic ... the problem at each search node A is to find a .... no constraint programmer would use such a.

A particle formulation for treating differential diffusion in ...
viable alternative to RANS in many areas, including combustion, and new ... Here we list the desirable properties of an ideal diffusion model ..... This is the extension of the Shvab–Zeldovich form of the energy equation [42] to unequal diffusiviti

Optimal Control Problems with State Specific Jumps in ...
Jan 22, 2013 - An important class of control problems in economics are those in which the state ... the state equation switches (jumps) each time the state variable ..... matching real world data would require adding more features to our model ...

Recommendations for in-situ data Near Real Time Quality ... - GitHub
data centre has some hope to be able to correct them in .... different from adjacent ones, is a spike in both size .... average values is more than 1°C then all.

The surprise element: jumps in interest rates
(1) Higher moment behavior: changes in interest rates demonstrate .... Thus reversals account for approximately 75% of the jump intensity. Thus, there is strong ...

Semi-supervised Learning in Nonstationary Environments
Author email: [email protected], and [email protected]. This work ...... learning in data streams," in Intl. Conf. on Discovery Sicence, 2006, pp. 29-40.

Diffusion In a Baggie.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Diffusion In a ...

Diffusion Characteristics for Simultaneous Source ...
coded signal, or from encrypted signal by making smaller changes in the key. .... [5] John G. Prakis, “Digital Communication”, Mc-Graw Hill Companies, (2007).

Extension Communication And Diffusion Of Innovations For ...
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Main menu.

Testing for Multiple Bubbles∗
Dec 23, 2011 - the test significantly improves discriminatory power and leads to distinct power gains when .... market exuberance arising from a variety of sources, including ... An alternative sequential implementation of the PWY procedure is.

Diffusion Adaptation Strategies for Distributed ... - IEEE Xplore
Oct 9, 2014 - data to estimate some. 1 parameter vector in a distributed manner. There are a ... recovery of sparse vectors to be pursued both recursively and.

Cycle Detection, Order Finding and Discrete Log with Jumps
unique solution modulo lcm(m1,...,mk) iff for all i, j, ai ≡ aj mod gcd(mi,mj). Lemma 4.3 ...... In Analytic Number Theory,. Proceedings of Symposia on Pure ...

Detecting Near-Duplicates for Web Crawling - Conferences
small-sized signature is computed over the set of shin- .... Another technique for computing signatures over .... detection mechanisms for digital documents.

Finding Near Duplicates in Short Text Messages in ...
Scalability is one of the main concerns of our project. Our implementation is able to scale up for ... repetitive tweets generated by third party apps (such as games, mobile advertisement app) and human spamming. ..... We are planning to applying our

Practical Significance Testing for Experiments in ...
Aug 12, 2016 - Data is correctly split into training, dev, test data. Systems A and B produce outputs oA and oB (on test data). Evaluation metric e gives .... 2 ∗ Pr ∗ Rec. Pr + Rec. Benjamin Roth (CIS LMU München). Practical Significance Testin

Testing for Violations of Microscopic Reversibility in ...
CA) was cotransfected as a marker with the cDNA of interest using the FuGENE .... likelihood fitting of each file (∼2.1 × 105 events) is 6.5 ( 0.1. (N ) 5, S.D.) with ...