1. Introduction
Precipitating cumulus convection that extends throughout the troposphere prevails in most regions in the Tropics. It has been known to exhibit various spatial and temporal scales tied to tropical waves and disturbances (e.g., Chang 1970). Information about cloud-top temperature deduced from satellite imagery, such as outgoing longwave radiation (OLR), has been used as an index of deep convective activity in the Tropics. Takayabu (1994) and Wheeler and Kiladis (1999) performed wavenumber–frequency spectral analyses on the satellite imagery and found wave signals corresponding to the equatorially trapped waves (e.g., Matsuno 1966) in their theoretical dispersion relations. These results indicate that tropical deep convection is, to a certain degree, coupled with tropical wave motion. Wheeler and Kiladis (1999) also pointed out exceptional phenomena such as the tropical depression–type disturbances (TD; Takayabu and Nitta 1993) and the 40–60-day Madden– Julian oscillation (MJO; Madden and Julian 1971), which do not have theoretical counterparts in the wavenumber–frequency domain. The authors of both papers noted strong spectral broadening in their analyses. Salby and Hendon (1994) found similar broadening in the OLR power spectra in the frequency domain over the Indian Ocean–western Pacific sector (i.e., the warm pool), and commented that the background variance cannot be accounted for by simple white noise or first-order red noise processes.
The broadening of power spectra in addition to the MJO and TD suggest a very high degree of complexity in the convectively coupled tropical motions. One manifestation of the complexity is the multiscale (“hierarchical”) structure of tropical convection (Nakazawa 1988). Convective organization over the warm pool appears to have spatial sizes from O(102) to O(103) km and temporal scales from O(1) to O(10) days. In other words, the observed organizations range from mesoscale convective systems (MCSs) following the course of the diurnal cycle to organized supercloud clusters embedded in the eastward-propagating planetary-scale MJO. It is known that different scales of organizations tend to interlock (e.g., Nakazawa 1988; Lau et al. 1991; Weng and Lau 1994; Tung and Yanai 2002b). Mapes (1993) attributes such behavior to the existing MCSs laying out a favorable environment for subsequent MCSs through gravity wave dynamics. In a statistical description, the variance of the convection-coupled dynamic system spans a wide range of scales. Furthermore, it is spread in the self-similar manner as observed by satellite imagery across this temporal or spatial range, as suggested by Yano and Takeuchi (1987) and Yano and Nishi (1989).
Fractal and multifractal concepts provide a systemic description of complex dynamic phenomena. For instance, they have been developed to understand turbulence phenomena (Frisch 1995). Convection is one type of turbulent motion in the atmosphere. It differs from classic turbulence because of the effects of latent heating, whereas the latter is dominated by dissipative processes. Many quantities related to convective processes have fractal or multifractal characteristics, for example, the cloud liquid water content (Davis et al. 1996; multifractal), convective available potential energy (CAPE; Yano et al. 2001; fractal), and convective momentum transport (Tung and Yanai 2002a; fractal). Yano et al. (1996) examined the dependence of the fractal feature of convective organization on various cumulus parameterization methods. Using an OLR dataset with high spatial resolution, Pierrehumbert (1996) concluded that the spatial variability of the high cloud field over the tropical Pacific is multifractal. In particular, fractals and multifractals have been widely applied to the analysis and modeling of rainfall fields (e.g., Marsan et al. 1996; Over and Gupta 1996; Perica and Foufoula-Georgiou 1996; Veneziano et al. 1996; Schmitt et al. 1998), focusing on the rainfall in the mature stage of storm systems.
The goals of this work are to examine and describe the multiscale temporal variability of deep convective systems in the Tropics using the multifractal approach. The nature of this study is different from that concerning the multifractal aspects of conventional turbulence or the study of single storm systems. It involves the collective behavior and interaction of many multiscale convective systems coupled by dynamics field in the tropical environment. The deep convection index (
2. The deep convection index
During the TOGA COARE IOP, the center of the deep convective activity is located slightly south of the equator. Figure 1 shows the time series of the hourly
In order to study the impact of spatial smoothing on the multifractal feature of the index, seven time series of
3. The multiplicative cascade process and the concept of multifractals
a. The multiplicative cascade model
The concepts of fractals and multifractals were first introduced by Mandelbrot (e.g., Mandelbrot 1974), and further developed to deepen the understanding of the intermittent turbulent phenomena (see Frisch 1995). The mathematical foundation of multifractal theories was provided by Hentschel and Procaccia (1983), Frisch and Parisi (1985), and Halsey et al. (1986), etc. Readers are referred to Gouyet (1996) for an excellent introduction.
In essence, multifractals are characterized by many or infinitely many power-law relations. To migrate the notion of a fractal set to one that manifests multifractal measures, one may partition the fractal set into many (typically interwoven) subsets, each characterized by a power law. This can either be characterized by a spectrum of fractal dimensions (i.e., the generalized dimensions) with each dimension for one subset, as we will introduce shortly; or it can be characterized by singular measures, which, in concept, gives the relative “weight” of each subset. A fuller picture of multifractals can be established by constructing a multifractal field. Multifractals can be constructed by a multiplicative cascade process model. Since the
At stage 0, consider a unit length. Associate it with a unit weight.
Divide the unit length into two (say, left and right) segments of equal length.
Also, divide the weight into two fractions, r and 1 − r, and assign them to the left and right intervals, respectively. The parameter r, called the multiplier, is a random variable governed by a probability density function (pdf) P(r), 0 ≤ r ≤ 1.
Each new interval and its associated weight are further divided into two parts following the same rules.
A scheme is shown in Fig. 3. At stage i, the length of the interval is 2−i, and its associated multiplier is denoted as rij with j = 1, … , 2i. If the random multiplier distribution, P(r), is assumed to be symmetric about r = 0.5 and have successive moments μ1, μ2, … , then both rij and 1 − rij have marginal pdf P(r). At stage N, the weights (i.e., the measures) are denoted as {wn(N), n = 1, … , 2N}; each element is expressed as wn(N) = u1u2 ⋯ u1 ⋯ uN, where ul, l = 1, … , N, is either rij or 1 − rij, i = 1, … , N. Therefore, {ul, l ≥ 1} are independent identically distributed (iid) random variables with pdf P(r).
Note that the above scheme is sometimes called microcanonical multiplicative cascades, where the multipliers add up exactly to 1. One may require that the weights are preserved only in terms of their expected value. Such a scheme is sometimes called canonical multiplicative cascades, and can easily be constructed by generating two random variables, for example, ri1 and ri2, where E(ri1 + ri2) = 1, and a weight wi at stage i is split into two parts, wri1 and wri2. Multiplicative multifractals can even be nonconservative, as in Frisch (1995, p. 166). For our study, we shall employ the simplest conservative model, where weights simply add up to 1.
b. The definition of a multifractal field
The first few dimensions of the dimension spectrum are of special interest. For instance, D0 is the same as the fractal dimension α in (2), D1 is called the information dimension, and D2 is equivalent to the exponent of the power-law-decaying correlation function of the field (Hentschel and Procaccia 1983). When the measures are uniform or only weakly nonuniform (monofractal), Dq is independent of q and all Dq are reduced to the fractal dimension. This is equivalent to τ(q) being a linear function of q. When the measures are nonuniform (multifractal), Dq is a monotonically decaying function of q and τ(q) is a nonlinear function of q.
The above two properties or their similar forms are frequently used as the criteria to determine whether a dataset is multifractal (e.g., Davis et al. 1996; Ivanova and Ackerman 1999). However, the most strict definition of a multifractal field, as illustrated by the multiplicative cascade model, is that the multiplier distribution function P(r) is stage independent. Given a sufficiently long dataset, it is possible to estimate P(r) at different stages [e.g., Frederiksen et al. (1997), who used O(105) samples]. This criterion is preferred since it is more free of artifacts. However, such datasets are rare in the atmospheric sciences.
If each measure, wn(N), in Fig. 4 is interpreted as the total convective activity in a time interval of length 2−NT, where T is the total time period of interest (2880 h), then the multiplicative cascade process becomes a primitive stochastic model for a time series of convective activity. Note the similarity between the intermittent features depicted in Figs. 1 and 4.
4. Multifractal characteristics of the ITBB time series and implications
a. Multifractal characterizations
Figure 5a shows −log2Mq(ϵ) versus −log2ϵ for the 5° × 5° averaged
The second step toward characterizing the
In Figs. 5b,c, the broken curves depict the τ(q) and Dq functions obtained from the model time series (Fig. 4). In the model, the τ(q) behaves similarly to that of the observed
The
b. Consequence of spatial averaging
We now examine the impact of spatial averaging on the multifractal features of the
Figure 7 shows the generalized dimensions (Dq) of the seven domain-averaged time series, which are plotted as in Fig. 5c. The small error bars for Dq in the larger domains indicate that the scaling relations are more well defined there. Functions Dq of larger domain averages descend more slowly with q than those of smaller domain averages, particularly at smaller q values. This suggests that spatial averaging reduces the significance of the multifractal characteristics of the time series but does not eliminate it completely. The Dq for the single-point dataset appears to be horizontal with q ≥ 2, most likely the result of insufficient samples, particularly those of relatively extreme events, since the time series is most sparse. Such deficiency in sampling is also reflected by the large error bars for the Dq estimation. Furthermore, with the same q, larger-domain averages have bigger Dq. This may reflect that physical processes with better-defined long-range trends are dominant in the time series of larger-domain averaging. This aspect will be discussed along with the variance– time relations in section 5.
c. Consequence of temporal smoothing
A rather common picture of data analysts is: x(t) = y(t) + n(t), where y(t) is the true signal, n(t) is the measurement noise, often taken as white, and x(t) is the observed signal. One might expect x(t) to have a well-defined mean when averaged over a long enough period of time, while its standard deviation is insignificant when compared with the mean; or, by applying a moving average, seemingly “transient” noise could be removed. This picture is at least partly true, since measurement noise is always present. However, the averaging or smoothing tactic may not be as constructive as expected, especially when the true signal displays multifractal features. This is because the variations of y(t) over a broad range of time scales are the intrinsic features of such processes, as succinctly and precisely captured by the definition of multifractal in (6). More concretely, the variance and transient features of x(t) are driven from both y(t) and n(t). While the contribution from n(t) may be negligibly small when the average is taken over a long time interval, the contribution from y(t) can nevertheless be quite large. Often, the mean of the signal can be much smaller than its standard deviation. In some cases, the mean of x(t) may not even be well defined.
We can quantify how effective it is to temporally smooth the
Figures 9a–d show the results of daily and monthly nonoverlapping window averages for all four time series. Standard deviations are plotted along with the monthly averages. Evidently, the window averages become well-defined much sooner in the white noise process and shuffled
5. Long-range dependence
a. Long-range dependence in the variance–time relation
In this section, we introduce an interesting property called long-range dependence in the
Figure 11a plots a time series with length 2880 generated from such a process with a = 0.97, which is the lag-1 autocorrelation coefficient of the 5° × 5°
Figure 11b shows variance–time plots of the red noise processes from the AR model with a = 0 (white noise), 0.1, 0.5, 0.7, 0.9, 0.97, and 0.97 with only positive values, in addition to that of the 5° × 5°
The AR model is not entirely without merits. As indicated in Figs. 10 and 11b, the rate of decreasing variance versus increasing time interval m of the 5° × 5°
b. Properties associated with long-range dependence
Mandelbrot and Van Ness (1968) coined the term “fractional Brownian motions” (fBms) for an extended family of statistically self-similar Gaussian functions characterized by the Hurst exponent, H, 0 < H < 1. For ordinary Brownian motion (Bm, a nonstationary process), H = 1/2. White Gaussian noise is the derivative of the Bm; similarly, the fractional Gaussian noise (fGn) is the derivative of the fBms. The fBm family is often called the random-walk-type process, while the fGn is called the increment process. Both provide useful models for natural phenomena containing both macro- and microscales. The fGn is a much simplified prototype to address the power-law decaying autocorrelation function and power spectrum of the
Furthermore, when the variance–time relation Var[X(m)] ∼ m2H−2 holds, through the Weiner–Khinchin theorem (Monin and Yaglom 1975), the power spectral density (PSD) for the time series X decays in a power-law manner: PSDX(f) ∼ 1/(f2H−1), with f the frequency. The PSD for the time series Y = {Yn : n = 0, 1, 2, …}, with Yn =
Suppose the mean value of the increment process X = {Xi : i = 0, 1, 2, …} is μ(x). If 0 ≤ H < 1/2, then when Xj > μ(x), it will be more likely to have Xj+1 < μ(x). If one takes Xj − μ(x) as the jump at time j, a jump-up is more likely followed by a jump-down when 0 ≤ H < 1/2. This is called the antipersistent correlation. The integration of X leads to a random-walk-type process (fBm) that is less nonstationary than the Bm. Conversely, if 1/2 < H < 1, then the times series X has the persistent correlation, or long-range dependence: a jump-up is more likely followed by another jump-up at the next time step, so is a jump-down followed by another jump-down. The integration X generates a random-walk-type process that is more nonstationary than the Bm.
Since 1/2 < H ≤ 1 for all seven
In Fig. 10, the shuffled 5° × 5°
The pdf of
6. Variation of deep convective intensity in the MJO
Thus far the time series examined represent the local (Eulerian) time change of deep convective activity. An intriguing question is how deep convection evolves inside a large-scale coherent dynamic system such as the MJO. Therefore, in this section, multifractal analysis is applied Lagrangian-wise to time series of
Figure 13 shows the time series of deep convective activity moving along the front (dashed line), center (solid line), and rear (dotted line) parts of the MJO obtained from (a) the 0.5 and (b) the 0.8 thresholds. Evidently, the threshold selections do not yield qualitatively different results. Both Figs. 13a,b show a “shift of regime” in terms of convective intensity around 18 December, as the major part of the MJO event moves past 150°E. After 18 December, the convection time series becomes more nonstationary with increasing intensity, possibly associated to the property of long-range dependence observed in section 5. The regime shift is more obvious in the front and center parts of the MJO envelope than in the rear.
Using (6), the partition functions [Mq(ϵ)] for these time series are calculated with ϵ = 2−N, N = 1, 2, … , 9. Figure 14a shows the log–log plots of Mq(ϵ) derived from the
Figure 14c shows the Dq of time series obtained with the 0.8 threshold. The Dq values in this figure are almost identical to those in Fig. 14b. The dramatic difference between the fractality of the convective intensity in the center of the MJO event and that on its edges is again observed. The fact that the time series on the edges have more significant multifractal features indicates that convective intensity there must be characterized by intermittency. It is known that the cloud clusters embedded in the MJO event often exhibit a burstlike behavior. As shown in Fig. 15, at the map time of 1200 UTC 20 December, maxima of the convective system observed are centered around 160°E. A burst occurs 12 h later, as shown in the following map. Even though the center of maximum convection remains around 160°E with a comparable overall strength, small-scale convective systems are intensified on the peripheral. Such a burst is very likely to be the excitation of smaller-scale waves and disturbances interlocked with the MJO event.
It is interesting to note that even though Fig. 15 or a time–longitude section illustrates that convective intensity varies more on the edges than in the center axis of the MJO, the multifractal analysis in Figs. 13 and 14 further reveals 1) the disturbances and waves embedded in the eastward-moving MJO are interdependent in the range of 1 h to about 5–10 days, characterized by power laws; and 2) the convection on the edges is more intermittent than that in the center and the interdependence of burst events on the edges can be further characterized with power laws.
7. Summary and concluding remarks
In this paper, we examined multiscale tropical deep convective variability in terms of the 4-monthlong hourly high-resolution deep convection index (
Although multifractal features exist in all the time series with different degrees of spatial averaging, the larger the spatial averaging, the less nonlinear the Dq function becomes. This indicates that multifractal features can be gradually weakened through area smoothing. Limited by the spatial coverage of the current dataset, the critical domain size for multifractal features to disappear cannot be addressed. Yano et al. (2001) used point-by-point station data during the TOGA COARE IOP and found that convective activities vary as 1/f stochastic processes. The results of our spatial-averaging analysis verify their conjecture that the 1/f property would remain significant after large-domain averaging.
It is also found that the
The evolution of cloud clusters inside the envelope of the first MJO event during the IOP is studied from a Lagrangian standpoint. The center axis of the MJO envelope is determined using two different thresholds of retaining the top 20% and 10% of the deep convective activity. However, the results are not sensitive to the thresholds. Time series along the front, center, and rear of the MJO event show that convection intensifies significantly when approaching the date line. The continuous intensification may be associated with the property of long-range dependence observed in the Eulerian experiments. It is also found that in the range of 1 h to about 5–10 days, the occurrence of tropical disturbances and waves embedded in this MJO are interpendent, characterized by power laws. The convection on the front and rear edges of the MJO event is more intermittent than that in the center. Description of the intermittency, observed as bursting cloud clusters, requires the use of higher-order partition functions and generalized dimensions.
Since convection is a spatiotemporal phenomenon, a more desirable framework to describe convective systems is to characterize their spatiotemporal multifractal properties. The description may start from first characterizing the spatial (temporal) multifractal features of the systems at a fixed time (space), then examining how the features vary with time (space). In more favorable situations, one may obtain some observables, which contain information both on the spatial and temporal complexities of the systems. One can then study the multifractal properties of the systems through those observables. For example, in the study of spatiotemporal chaos through coupled map approaches, a commonly used observable is the mean of the variables on the lattices. That is, one proceeds by calculating dimension densities of this observable (e.g., Bauer et al. 1993; Politi and Witt 1999; Raab and Kurths 2001). However, understanding using this path has been limited. Due to the limited spatial sample size of the data studied here, we have mostly performed temporal multifractal analysis, together with spatial averaging. It is interesting to note that with larger domain spatial averaging, the Dq curves become larger (Fig. 7). Qualitatively this is consistent with the concept of dimension density in the study of spatiotemporal chaos. However, we emphasize that geophysical systems are far more complex than simple spatiotemporal chaotic systems modeled by coupled map lattices. Furthermore, due to the poor spatial sampling of the data, it is impossible to practice the concept of dimension density. At this juncture, we draw readers' attention to the recent work by Pierrehumbert (1996), who used a dataset with 3 times the spatial sample size of ours to examine the spatial multifractal properties of the high cloud variability over the tropical Pacific Ocean.
The multifractal features of the 5° × 5°
The same kind of multiplicative processes that induce multiscale convective variability in the Tropics may be also present in other convective regions on the earth. However, in the Tropics, the scale-under influence of a process could be maximized by the weak Coriolis force and relatively uniform large-scale thermodynamic variability. As suggested by our study, the multifractal features could be preserved beyond the spatial size of 25° × 25° and so could the maximum scale of influence. It is reasonable to speculate that such maximum scale is constrained by the latitudinal extent of the warm pool.
The multifractal approach used here has many other applications. It is a potent tool to evaluate model performance, reaching beyond mean and variance to provide information in higher-order statistical moments that are associated with the correlation structure of extreme events. For instance, with longer data records of convective activity, the scaling region (currently from 1 h to about 20 days) could be reexamined and may be extended, particularly to test whether the wide-range variability of MJO in observation can also be described with power laws. The results would be a more quantitative evaluation of current models' ability to produce MJO along with its well-known irregularity and interlocking scales. Finally, a most complex problem in modeling a convecting atmosphere, notably the Tropics, is characterizing interactions among scales. The analysis herein shows that extreme events play an important role in the observed convective variability. From the perspective that cumulus parameterization should represent extreme events and large-scale models such as GCMs are sensitive to the parameterization, subgrid-scale extreme convective events likely have identifiable effects at large scales. The multifractal approach uses few parameters to describe not only the occurrence of the extreme events but also the interdependence of these events: a parsimonious evaluation of cumulus parameterization. The multifractal technique could be used to analyze cloud-system-resolving model datasets of large dynamic range, such as explicit simulations of convectively coupled tropical waves, in order to gain insight into the attendant multiscale interactions.
Acknowledgments
The authors are very grateful to discussions with D. Baker, P. Bechtold, H. Hastings, C. Jeffery, G. Kiladis, J. Lin, C. Liu, H. Liu, M. Montgomery, and R. Plougonven; as well as invaluable suggestions from B. Mapes, J.-I. Yano, and two anonymous reviewers. The high-resolution deep convective index is the courtesy of T. Nakazawa. The leading author also deeply appreciates M. Yanai for stimulating debates while the work began to materialize. This work is supported by the NCAR Advanced Study Program (ASP).
REFERENCES
Bauer, M., H. Heng, and W. Martienssen, 1993: Characterization of spatiotemporal chaos from time-series. Phys. Rev. Lett, 71 , 521–524.
Biferale, L., M. Blank, and U. Frisch, 1994: Chaotic cascades with Kolmogorov 1941 scaling. J. Stat. Phys, 75 , 781–795.
Box, G. E. P., and G. M. Jenkins, 1976: Time Series Analysis: Forecasting and Control. 2d ed. Holden-Day, 553 pp.
Chang, C-P., 1970: Westward propagating cloud patterns in the tropical Pacific as seen from time-composite satellite photographs. J. Atmos. Sci, 27 , 133–138.
Davis, A. B., A. L. Marshak, W. J. Wiscombe, and R. F. Cahalan, 1996: Multifractal characterizations of intermittency in nonstationary geophysical signals and fields—A model-based perspective on ergodicity issues illustrated with cloud data. Proc. Second Workshop on Nonstationary Random Processes and Their Applications, San Diego, CA, World Scientific, 97–158.
Frederiksen, R. D., W. J. A. Dahm, and D. R. Dowling, 1997: Experimental assessment of fractal scale similarity in turbulent flows. Part 3. Multifractal scaling. J. Fluid Mech, 338 , 127–155.
Frisch, U., 1995: Turbulence—The legacy of A. N. Kolmogorov. Cambridge University Press, 296 pp.
Frisch, U., and G. Parisi, 1985: Fully developed turbulence and intermittency. Turbulence and Predictability in Geophysical Fluid Dynamics and Climate Dynamics (International School of Physics Énrico Fermi, Course 88), M. Ghil, R. Benzi, and G. Parisi, Eds., North-Holland, 71–88.
Gao, J. B., and I. Rubin, 2001: Multiplicative multifractal modelling of long-range-dependent network traffic. Int. J. Comm. Syst, 14 , 783–801.
Gouyet, J-F., 1996: Physics and Fractal Structures. Springer-Verlag, 234 pp.
Halsey, T. C., M. H. Jensen, L. P. Kadanoff, I. Procaccia, and B. Shraiman, 1986: Fractal measures and their singularities: The characterization of strange sets. Phys. Rev. A, 33 , 1141–1151.
Hayashi, V., and D. G. Golder, 1985: Nonlinear energy transfer between stationary and transient waves simulated by a GFDL spectral general circulation model. J. Atmos. Sci, 42 , 1340–1345.
Hentschel, H. G. E., and I. Procaccia, 1983: The infinite number of generalized dimensions of fractals and strange attractors. Physica D, 8 , 435–444.
Ivanova, K., and T. Ackerman, 1999: Multifractal characterization of liquid water in clouds. Phys. Rev, 59E , 2778–2782.
Kaulakys, B., and T. Meskauskas, 1998: Modeling 1/f noise. Phys. Rev. E, 58 , 7013–7019.
Kolmogorov, A. N., 1941: The local structure of turbulence in incompressible viscous fluid for very large Reynolds number. Dokl. Akad. Nauk SSSR, 30 , 299–303.
Lau, K-M., T. Nakazawa, and C. H. Sui, 1991: Observations of cloud cluster hierarchies over the tropical western Pacific. J. Geophys. Res, 96 , 3197–3208.
Lilly, D. K., 1983: Stratified turbulence and the mesoscale variability of the atmosphere. J. Atmos. Sci, 40 , 749–761.
Lin, J. W-B., and J. D. Neelin, 2002: Considerations for stochastic convective parameterization. J. Atmos. Sci, 59 , 2793–2823.
Madden, R. A., and P. R. Julian, 1971: Detection of a 40–50-day oscillation in the zonal wind in the tropical Pacific. J. Atmos. Sci, 28 , 702–708.
Mallat, S. G., 1999: A Wavelet Tour of Signal Processing. 2d ed. Academic Press, 637 pp.
Mandelbrot, B. B., 1974: Intermittent turbulence in self-similar cascades: Divergence of high moments and dimension of the carrier. J. Fluid Mech, 62 , 331–358.
Mandelbrot, B. B., 1982: The Fractal Geometry of Nature. Freeman, 468 pp.
Mandelbrot, B. B., and J. W. Van Ness, 1968: Fractional Brownian motions, fractional noises and applications. SIAM Rev, 10 , 422–437.
Mapes, B. E., 1993: Gregarious tropical convection. J. Atmos. Sci, 50 , 2026–2037.
Marsan, D., D. Schertzer, and S. Lovejoy, 1996: Causal space–time multifractal processes: Predictability and forcasting of rain fields. J. Geophys. Res, 101 , 26333–26346.
Matsuno, T., 1966: Quasi-geostrophic motions in the equatorial area. J. Meteor. Soc. Japan, 44 , 25–43.
Monin, A. S., and A. M. Yaglom, 1975: Statistical Fluid Mechanics. Vol. 2. MIT Press, 874 pp.
Murakami, M., 1983: Analysis of the deep convective activity over the Western Pacific and Southeast Asia. Part I: Diurnal variation. J. Meteor. Soc. Japan, 61 , 60–76.
Nakazawa, T., 1988: Tropical super clusters within intraseasonal variations over the western Pacific. J. Meteor. Soc. Japan, 66 , 823–839.
Nakazawa, T., 1995: Intraseasonal oscillations during the TOGA-COARE IOP. J. Meteor. Soc. Japan, 73 , 305–319.
Over, T. M., and V. K. Gupta, 1996: A space–time theory of mesoscale rainfall using random cascades. J. Geophys. Res, 101 , 26319–26331.
Perica, S., and E. Foufoula-Georgiou, 1996: Model for multiscale disaggregation of spatial rainfall based on coupling meteorological and scaling descriptions. J. Geophys. Res, 101 , 26347–26361.
Pierrehumbert, R., 1996: Anomalous scaling of high cloud variability in the tropical Pacific. Geophys. Res. Lett, 23 , 1095–1098.
Politi, A., and A. Witt, 1999: Fractal dimension of space–time chaos. Phys. Rev. Lett, 82 , 3034–3037.
Raab, C., and J. Kurths, 2001: Estimation of large-scale dimension densities. Phys. Rev. E, 64 , 016216.
Salby, M. L., and H. H. Hendon, 1994: Intraseasonal behavior of clouds, temperature, and motion in the Tropics. J. Atmos. Sci, 51 , 2207–2224.
Schmitt, F., S. Vannitsem, and A. Barbosa, 1998: Modeling of rainfall time series using two-state renewal processes and multifractals. J. Geophys. Res, 103 , 23181–23193.
Shapiro, L. J., and D. E. Stevens, 1980: Parameterization of convective effects on the momentum and vorticity budgets of synoptic-scale Atlantic tropical waves. Mon. Wea. Rev, 108 , 1816–1826.
Sui, C. H., and M. Yanai, 1986: Cumulus ensemble effects on the large-scale vorticity and momentum fields of GATE. Part I: Observational evidence. J. Atmos. Sci, 43 , 1618–1642.
Takayabu, Y. N., 1994: Large-scale cloud disturbances associated with equatorial waves. Part I: Spectral features of the cloud disturbances. J. Meteor. Soc. Japan, 72 , 433–449.
Takayabu, Y. N., and T. Nitta, 1993: 3–5 day period disturbances coupled with convection over the tropical Pacific Ocean. J. Meteor. Soc. Japan, 71 , 221–246.
Tollerud, E. I., and S. K. Esbensen, 1983: An observational study of the upper-tropospheric vorticity fields in GATE cloud clusters. Mon. Wea. Rev, 111 , 2161–2175.
Tung, W-W., and M. Yanai, 2002a: Convective momentum transport observed during the TOGA COARE IOP. Part I: General features. J. Atmos. Sci, 59 , 1857–1871.
Tung, W-W., and M. Yanai, 2002b: Convective momentum transport observed during the TOGA COARE IOP. Part II: Case studies. J. Atmos. Sci, 59 , 2535–2549.
Veneziano, D., R. L. Bras, and J. D. Niemann, 1996: Nonlinearity and self-similarity of rainfall in time and a stochastic model. J. Geophys. Res, 101 , 26371–26392.
Webster, P. J., and R. Lukas, 1992: TOGA COARE: The coupled ocean–atmosphere response experiment. Bull. Amer. Meteor. Soc, 73 , 1377–1416.
Weng, H., and K-M. Lau, 1994: Wavelets, period doubling, and time-frequency localization with application to organization of convection over the tropical western Pacific. J. Atmos. Sci, 51 , 2523–2541.
Wheeler, M., and G. N. Kiladis, 1999: Convectively coupled equatorial waves: Analysis of clouds and temperature in the wavenumber-frequency domain. J. Atmos. Sci, 56 , 374–399.
Wu, X., and M. Yanai, 1994: Effects of vertical wind shear on the cumulus transport of momentum: Observations and parameterization. J. Atmos. Sci, 51 , 1640–1660.
Yanai, M., B. Chen, and W-W. Tung, 2000: The Madden–Julian oscillation observed during the TOGA COARE IOP: Global view. J. Atmos. Sci, 57 , 2374–2396.
Yano, J-I., and Y. Takeuchi, 1987: The self-similarity of horizontal cloud pattern in the intertropical convergence zone. J. Meteor. Soc. Japan, 65 , 661–667.
Yano, J-I., and N. Nishi, 1989: The hierarchy and self-affinity of the time variability within the tropical atmosphere inferred from the NOAA OLR data. J. Meteor. Soc. Japan, 67 , 771–789.
Yano, J-I., J. C. McWilliams, and M. W. Moncrieff, 1996: Fractality in idealized simulations of large-scale tropical cloud systems. Mon. Wea. Rev, 124 , 838–848.
Yano, J-I., K. Fraedrich, and R. Blender, 2001: Tropical convective variability as 1/f noise. J. Climate, 14 , 3608–3616.
The National Center for Atmospheric Research is sponsored by the National Science Foundation.