<<

456 WEATHER AND FORECASTING VOLUME 24

Objective Estimation of the 24-h Probability of Formation

ANDREA B. SCHUMACHER Cooperative Institute for Research in the Atmosphere–Colorado State University, Fort Collins, Colorado

MARK DEMARIA AND JOHN A. KNAFF NOAA/NESDIS/Center for Satellite Applications and Research, Fort Collins, Colorado

(Manuscript received 31 December 2007, in final form 28 May 2008)

ABSTRACT

A new product for estimating the 24-h probability of TC formation in individual 58358 subregions of the North Atlantic, eastern North Pacific, and western North Pacific tropical basins is developed. This product uses environmental and convective parameters computed from best-track tropical cyclone (TC) positions, National Centers for Environmental Prediction (NCEP) Global Forecasting System (GFS) analysis fields, and water vapor (;6.7 mm wavelength) imagery from multiple geostationary satellite platforms. The pa- rameters are used in a two-step algorithm applied to the developmental dataset. First, a screening step removes all data points with environmental conditions highly unfavorable to TC formation. Then, a linear discriminant analysis (LDA) is applied to the screened dataset. A probabilistic prediction scheme for TC formation is developed from the results of the LDA. Coefficients computed by the LDA show that the largest contributors to TC formation probability are climatology, 850-hPa circulation, and distance to an existing TC. The product was evaluated by its Brier and relative operating characteristic skill scores and reliability diagrams. These measures show that the algorithm- generated probabilistic forecasts are skillful with respect to climatology, and that there is relatively good agreement between forecast probabilities and observed frequencies. As such, this prediction scheme has been implemented as an operational product called the National Environmental Satellite, Data, and In- formation Services (NESDIS) Tropical Cyclone Formation Probability (TCFP) product. The TCFP product updates every 6 h and displays plots of TC formation probability and input parameter values on its Web site. At present, the TCFP provides real-time, objective TC formation guidance used by tropical cyclone forecast offices in the Atlantic, eastern Pacific, and western Pacific basins.

1. Introduction such as the Joint Warning Center and the Central Pacific Hurricane Center have similar require- Beginning in 2003, the National Hurricane Center ments. Although subjectively generated, these forecasts (NHC) extended their official tropical cyclone (TC) track rely heavily on various forms of objective guidance. and intensity forecasts from 3 to 5 days, based upon the At this time, the most widely used source for TC need for longer lead times. As forecasts are extended, formation guidance is global numerical models. As the it becomes increasingly possible for storms to form after resolution of global models increases, so does their the start of and have an impact before the end of promise in predicting TC formation (Pasch et al. 2006; a forecast period, intensifying the need for TC forma- Harr 2007). Yet uncertainty regarding their demon- tion guidance. Currently, the NHC has an operational strated forecast skill and model-based biases, including requirement to provide TC formation forecasts in its the tendency of these models to overpredict TC for- tropical weather outlooks. Other TC forecast centers mation (Beven 1999), still limits their utility in TC forecasting. For these reasons, new objective methods Corresponding author address: Andrea Schumacher, CIRA– Colorado State University, Foothills Campus Delivery 1375, Fort of estimating TC formation probability are still desired. Collins, CO 80523-1375. Tropical cyclone formation is a rare event, and hence E-mail: [email protected] forecasting its likelihood on subbasin scales is a

DOI: 10.1175/2008WAF2007109.1

Ó 2009 American Meteorological Society Unauthenticated | Downloaded 10/06/21 01:21 PM UTC APRIL 2009 S C H U M A C H E R E T A L . 457

FIG. 1. The three basins covered by the TC formation product: the western North Pacific, eastern–central North Pacific, and Atlantic.

challenging task. To illustrate this point, consider the tive vortices in the presence of deep convective bursts Atlantic tropical cyclone basin as defined in Fig. 1. (McBride and Zehr 1981; Zehr 1992), and convective- There are 148 subregions of 58358 latitude–longitude scale features such as individual ‘‘hot towers’’ (Simpson within this basin and 183 days in the official Atlantic et al. 1998) and their associated thermodynamic and hurricane season (1 June–30 November). During the dynamic anomalies [e.g., ‘‘vortical hot towers’’; Hendricks period of 1949–2005, there was an average of 10.5 for- et al. (2004); Montgomery et al. (2006); Hendricks and mations per year over the Atlantic basin. Thus, in any Montgomery (2006)] . given 58358 subregion, there is less than a 1 in 2579 Many of these findings have been incorporated into chance (0.039%) of a TC forming within any given 24-h parameters for predicting TC formation. Gray (1975) period. A similar analysis performed for the eastern and created a seasonal genesis frequency parameter for the western North Pacific basins yields TC formation western North Pacific basin that was defined by the probabilities of 0.051% and 0.101%, respectively. multiplicative combination of six primary genesis pa- This ‘‘needle in a haystack’’ problem has received rameters: vertical shear, Coriolis parameter, low-level much attention over the years, and identifying and un- vorticity, sea surface temperature, moist stability, and derstanding the environmental conditions necessary for midlevel relative humidity. This work was later ex- TC formation has been the topic of numerous studies. panded to a global framework (Gray 1979). The cli- For example, Riehl (1948) studied TC formation in the matological values of this genesis frequency parameter North Pacific and found the upper-level disturbances in were shown to be well correlated with observed TC the vicinity of the low-level trades tended to enhance formation (McBride 1995). Expanding on this ap- low-level instability, which can lead to TC formation. proach, Emanuel and Nolan (2004) created a similar The global observational study of Gray (1968) noted index that was more applicable to global climate model that TCs tend to form in regions of weak vertical shear output. Similarly, DeMaria et al. (2001) developed a of the horizontal winds, which has been confirmed tropical cyclone genesis parameter (GP) for the tropical by others (e.g., Gray 1984; McBride and Zehr 1981; Atlantic basin that relied on 5-day running means of Bracken and Bosart 2000). Palme´n (1948) discussed the vertical shear, vertical instability, and midlevel mois- link between TC formation and sea surface temperature ture. They found that variations in the individual pa- (SST) and concluded that TCs typically form in regions rameter inputs and the subsequent changes in GP where the SST exceeds 268C. It has been shown that dry helped to explain intra- and interseasonal variations in air at midlevels, via its entrainment into convective TC formation. updraft cores and subsequent reduction in both buoy- Perrone and Lowe (1986, hereafter PL86) and ancy and precipitation efficiency (Ruprecht and Gray Hennon and Hobgood (2003, hereafter HH03) took a 1974; Gray 1975; Bister and Emanuel 1997), can inhibit more focused approach by developing TC formation sustained deep convection and have a negative effect on prediction schemes for use with tropical cloud clusters. TC formation. Furthermore, the explicit role of tropical Both studies developed algorithms using a statistical convection in TC formation has received considerable technique known as linear discriminant analysis (LDA) attention. Research in this area has spanned various to determine which tropical cloud clusters would de- spatial scales, including large-scale factors such as con- velop into tropical cyclones. PL86 used Navy Fleet vective instability (Gray 1975; Ooyama 1990; DeMaria Numerical Oceanography Central (NFNOC) analyzed et al. 2001) and the intertropical convergence zone archive fields and a step-wise discriminant analysis (ITCZ), the development of TCs from mesoscale dis- routine to develop a statistical forecast algorithm turbances such as easterly waves and midlevel convec- for TC genesis in the western North Pacific, and found

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC 458 WEATHER AND FORECASTING VOLUME 24 that the algorithm possessed skill when compared to domain is broken down into 58358 latitude–longitude climatology. The algorithm developed in HH03 ana- subregions,1 which divide the domain into a set of lyzed a 3-yr sample (1998–2000) of developing and 50 3 9 5 450 grid boxes. Environmental and convective nondeveloping cloud clusters in the tropical North At- parameters and formation probabilities are calculated lantic that were subjectively identified from geostationary over each of these subregions every 6 h at 0000, 0600, satellite infrared imagery. HH03 used environmental data 1200, and 1800 UTC. from the National Centers for Environmental Prediction– Information regarding historical tropical cyclone National Center for Atmospheric Research (NCEP– formation frequencies and locations was obtained from NCAR) reanalysis fields (Kalnay et al. 1996) and applied archived best-track files supplied by the NHC, CPHC, a discriminant analysis muchlikethatusedbyPL86.Like and JTWC. Each basin’s best tracks from 1949 to 2005 PL86, they too found that their discriminant analysis were merged and reconciled for duplicate storm entries. technique displayed skill at predicting TC formation from For the purpose of this study, TC formation (or genesis) preexisting tropical cloud clusters. was defined as the first time and position in the best In this paper, the approaches of PL86 and HH03 are tracks when a storm was classified as a tropical system. generalized to develop a tropical cyclone formation System types included in the archived best tracks that probability scheme for real-time use over multiple are deemed tropical under this classification scheme tropical basins. A linear discriminant analysis is applied include tropical depressions, tropical storms, and to the NCEP Global Forecasting System (GFS) model hurricanes/, while subtropical storms and analysis fields and water vapor (6.7 mm) imagery from extratropical storms are excluded. Track and intensity several geostationary satellites covering the North At- information for unnamed tropical depressions are not lantic, eastern North Pacific, and western North Pacific recorded in the official best tracks. For this reason, . However, unlike PL86 and HH03, information on unnamed tropical depressions was this study uses convective parameters in combination obtained from the Automated Tropical Cyclone Fore- with environmental parameters in lieu of limiting an- cast System (ATCF; Sampson and Schrader 2000). The alysis to existing tropical cloud clusters. Removing the ATCF was implemented in 1988, so unnamed tropical need for subjective identification of cloud clusters all- depression tracks were included from 1989 to 2005. The ows for the development of an objective, spatially and locations of all TC genesis cases included in the analysis temporally continuous product for estimating TC for- dataset are shown in Fig. 2. mation probability over an extended analysis domain. Monthly climatological TC formation probabilities Section 2 provides a description of the datasets used were calculated by adding up the total number of TC in this study. Section 3 describes the algorithm devel- formations in each subregion for each month over the opment and section 4 provides a statistical evaluation of 1949–2005 dataset and then dividing by the number of the algorithm over dependent and independent data- years (57). Figure 3 shows monthly climatological TC sets. The operational product that has been developed formation probabilities over the analysis domain for the and implemented from the algorithm is described in month of September. The largest monthly TC formation section 5. Section 6 summarizes the results of this study probability occurs in September between 158–208N and and discusses plans for future work. 1058–1108W in the eastern Pacific basin (red grid box in Fig. 3). This value represents 31 TC formations that have occurred in that subregion during the month of 2. Datasets September from 1949 to 2005, giving a monthly forma- The purpose of this study is to provide a new objective tion probability of 54.4% and hence an average daily product for estimating the probability of tropical cy- formation probability of 1.8%. This climatological TC clone formation in the National Hurricane Center formation probability is relatively small, especially for (NHC), Central Pacific Hurricane Center (CPHC), and the region with the highest number of TC formations Joint Typhoon Warning Center (JTWC) regions of per unit area in the world, which once again demon- forecast responsibility. The analysis domain extends strates the needle-in-a-haystack nature of predicting TC from the equator to 458N and from 1008Eto108W, formation. which covers all of the NHC and CPHC areas of re- Environmental conditions within each subregion sponsibility and a portion of the JTWC area of respon- were determined using GFS analysis and reanalysis sibility. The analysis domain consists of three basins: data. Prior to 2001, the NCAR–NCEP reanalysis fields, the Atlantic, eastern Pacific and western Pacific, whose boundaries are determined by geostationary satellite 1 Hereafter, any reference to ‘‘subregions’’ is referring this de- coverage and agency responsibility (Fig. 1). The analysis fined set of 58358 latitude–longitude grid boxes.

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC APRIL 2009 S C H U M A C H E R E T A L . 459

FIG. 2. All tropical cyclone formation points from the 1949–2005 best tracks. available on a 2.5832.58 global grid, were used. From imagery were available. The NCEP–NCAR model 2001 to 2005, the NCEP global operational analysis data reanalyses were available back to 1980, making satellite fields, available on a 18318 global grid, were used. Both data availability the limiting factor. Hence, the analysis datasets were interpolated to a uniform global 28328 time periods for the Atlantic, eastern Pacific, and grid. western Pacific basins extend from 1995, 1998, and 2000 In addition to the GFS data fields, water vapor (6.7 to 2005, respectively. mm) imagery from geostationary satellites was used to calculate parameters related to deep convective activity. Since the analysis domain is too large to be covered by a 3. Algorithm development single geostationary satellite, imagery from three cur- The algorithm developed here is similar to that of rently operational geostationary satellites is needed to Knaff et al. (2008, hereafter K08). The K08 procedure cover the analysis domain: Geostationary Operational involved two steps—a screening step and then a linear Environmental Satellite-E (GOES-E, centered at 758W), discriminant analysis step—that used environmental GOES-W (centered at 1358W), and Multifunctional values from satellite infrared imagery and analyses from Transport Satellites-1R (MTSAT-1R,centeredat1408E). the Statistical Hurricane Intensity Prediction Scheme Full-disk images from GOES-E and GOES-W were (SHIPS; DeMaria et al. 2005) model to identify a subset available back to 1995 and 1998, respectively, from the of tropical cyclones known as annular hurricanes. Since Cooperative Institute for Research in the Atmosphere the present study also seeks to discriminate between (CIRA) satellite archives and the National Oceanic and two groups—TC formation and nonformation cases— Atmospheric Administration (NOAA) Comprehensive based on environmental conditions, the same two-step Large Array-Data Stewardship System (CLASS). Imag- methodology is used. The first step involves screening ery from MTSAT-1R wasobtainedfromCIRAarchives out all sample set data points for which tropical cyclone dating back to November of 2005. Prior to November formation is highly unlikely. The second step involves 2005, the region currently covered by MTSAT-1R was the use of a linear discriminant analysis (see Wilks covered by GOES-9 (centered at 1558E) and Geosta- 2006), which allows one to distinguish between two tionary Meteorological Satellite-05 (GMS-05, centered at groups on the basis of a k-dimensional vector of ob- 1408E). Imagery from GOES-9 was obtained for April servations, x. In this case, x represents a list of envi- 2003–November 2005 from NOAA CLASS and imagery ronmental and convective parameters expected to have for GMS-05 dating back to 2000 was obtained from the relationships with TC formation. Tropical Regional and Mesoscale Branch K08 assesses the effectiveness of its algorithm by (RAMM) Advanced Meteorological Satellite Demon- evaluating hit rates and false alarm rates. Given the stration and Interpretation System (RAMSDIS) archives similarities of our analysis methods, we will use the 2 at CIRA. A summary of the geostationary satellite same strategy where applicable. These metrics rely on water vapor imagery sources collected for this study is four quantities that can be obtained from a standard shown in Fig. 4. All water vapor imagery was obtained 2 3 2 contingency table as described in Wilks (2006) from full-disk scans and was remapped to a 16-km and has been reproduced in Table 1. In this table, a is Mercator projection prior to analysis. The Levitus cli- the number of observed TC formation events that were matological monthly SSTs (Levitus 1982) were also used. forecasted as formation events by the algorithm (i.e., The only cases included in this analysis were those for ‘‘hits’’), b is the number of observed nonformation which both model reanalysis and satellite water vapor events that were forecasted as formation events by the algorithm (i.e., ‘‘false alarms’’), c is the number of ob- 2 Tropical RAMSDIS archives only included satellite imagery served formation events that were forecasted as non- from May to November. formation events by the algorithm (i.e., ‘‘misses’’), and d

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC 460 WEATHER AND FORECASTING VOLUME 24

FIG. 3. Climatological TC formation probabilities for the month of September, calculated from the 1949–2005 best tracks. is the number of observed nonformation events that shear (VSHEAR), 850-hPa circulation (CIRC), verti- were forecasted as nonformation events (i.e., ‘‘correct cal instability (THDEV), 850-hPa horizontal diver- negatives’’). Following this convention of Wilks (2006), gence (HDIV), latitude (LAT), percent land coverage the hit rate, which is also commonly referred to as the (PLAND), distance to any existing TC (DSTRM), probability of detection, is defined as HR 5 a/(a 1 c). climatological SST (CSST), cloud-cleared water vapor The false alarm rate is defined as FAR 5 b/(b 1 d). brightness temperature (BTWARM), percentage of cold pixel coverage (PCCOLD), and 24-h climatolog- a. Environmental parameters ical TC formation probability (CPROB). Additional information on parameter choices and methods of Before either step of the algorithm could begin, a set calculation is provided in the appendix. of environmental and convective parameters had to be chosen. As described in section 1, numerous environ- mental conditions have been found to have an effect on b. Screening step TC formation. Input parameters were taken from the The first step of the algorithm consists of screening more robust results of TC formation research, and only out all cases for which TC formation is highly unlikely. those values that were calculable on subregional scales Although the parameters chosen for this study influence from the available datasets were used. The input pa- TC formation in all regions of the analysis domain, the rameters chosen for this study are 850–200-hPa vertical relative magnitudes of these influences may vary from basin to basin. For this reason each basin was treated as an independent dataset during algorithm development. The same procedure, described here, was applied to each basin individually. To determine the appropriate screening thresholds, all cases in the developmental dataset were partitioned into two groups: TC genesis (TCG) and no TC genesis (NTCG). Here, a ‘‘case’’ refers to any given 58358 latitude–longitude subregion at a single analysis time of 0000, 0600, 1200, or 1800 UTC. When a tropical cyclone forms, all cases spatially coincident with the formation location at times within 24 h prior to the formation time are assigned to the TCG group. Since the analysis is

FIG. 4. Inventory of water vapor imagery from geostationary TABLE 1. The 2 3 2 contingency table values, as defined in satellites used in the analysis. The abscissa labels correspond to Wilks (2006). 1 January of each year (e.g., J-98 stands for 1 Jan 1998). The or- dinate labels are index values that correspond to the numbers Observed shown to the left of each geostationary satellite listed in the legend. Yes No Satellite indices 1–3 represent western Pacific satellite coverage, Forecast Yes a (hits) b (false alarms) 4–5 represent eastern Pacific satellite coverage, and 6–7 represent No c (misses) d (correct negatives) Atlantic satellite coverage.

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC APRIL 2009 S C H U M A C H E R E T A L . 461

TABLE 2. The criteria used to eliminate data points from the dataset during the screening step. Note that CPROB was not used for screening because it is derived from a limited best track dataset (1949–2005). In the third column, NCEP indicates the data source is the NCEP GFS analyses and SAT WV indicates the data source is satellite water vapor (6.7 mm) imagery.

Elimination criteria Abbreviation Screening parameter Data source Atlantic Eastern Pacific Western Pacific LAT Lat (8N) Domain definition ,5.0 ,5.0 ,5.0 PLAND Land coverage (%) Land mask $100.0 $100.0 $100.0 DSTRM Distance to nearest TC (8 lat–lon) Best tracks ,2.5 ,2.5 ,2.5 CSST Max climatological SST (8C) Levitus SST ,21.0 ,21.0 ,21.0 VSHEAR 200–850-hPa vertical shear (m s21) NCEP .25.2 .15.9 .19.5 CIRC 850-hPa circulation (m s21) NCEP ,21.5 ,21.2 ,20.9 THDEV Vertical instability (8C) NCEP ,22.6 ,23.0 ,1.6 HDIV 850-hPa horizontal divergence (31025 s21) NCEP .1.0 .0.7 .0.5 PCCOLD Cold pixel count (%) SAT WV ,2.8 ,5.0 ,3.0 BTWARM Avg cloud-cleared brightness temperature (8C) SAT WV .225.4 .223.1 .227.8 performed at 6-h intervals, each TC formation in the alone an ineffective predictor of TCG. For example, in best-track record represents four TCG cases in the de- the Atlantic there are 701 TCG cases and 1 818 983 velopmental dataset. All cases not assigned to the TCG NTCG cases in the development dataset. A hit rate of group were considered NTCG cases. 95% corresponds to 665 correctly identified TCG cases. The only parameter listed in section 3a not used for To compare, a false alarm rate of 24% corresponds to screening is CPROB. This value is computed from a 428 553 false alarms. This results in a TC formation finite (57 yr) record. As such, when a subregion has a probability of 100 3 (665/428 553) 5 0.16%. Although value of CPROB 5 0, it does not necessarily imply that screening results in a probability that is a fourfold im- tropical cyclone formation is highly unlikely, but rather provement over the 0.039% random probability com- indicates that a tropical cyclone has not occurred at that puted in section 1, the staggering number of false alarms particular place and time in the last 57 yr. The lack of a in relation to hits makes screening alone ineffective for physical relationship between CPROB and TC forma- the purposes of forecasting TC formation. tion along with the limited nature of the climatological c. Linear discriminant analysis step dataset make it unsuitable for use as a screening pa- rameter. The range of parameter values for each group To improve the skill of our algorithm over that pro- was examined, and screening criteria were defined so vided by screening alone, LDA (Wilks 2006) was that no more than 1% of TCG cases were removed by applied to the screened developmental dataset. This any one parameter’s screening criterion. With 10 screen- procedure uses the parameter values and group mem- ing criteria, this method guarantees that a maximum of bership (i.e., TCG or NTCG cases) of each case in the 10% of TCG cases would be eliminated from the analysis developmental dataset to solve for a set of coefficients dataset. The final criteria chosen (Table 2) eliminated that can be used to determine the group membership of approximately 5% of TCG cases in each basin while independent data points. In other words, for any inde-

76%, 89%, and 75% of NTCG cases were eliminated pendent data point with parameter values x1, x2, ..., xn, from the Atlantic, eastern Pacific, and western Pacific the discriminant function is datasets, respectively (Table 3, first two rows). n If we were to evaluate the screening step alone as a f 5 a0 1 åi51aixi, forecast, where all data points screened out are equiv- alent to a forecast of NTCG and all points retained are where n is the number of parameters and a0 is a constant considered TCG forecasts, we could compute the cor- term. This function can be used to determine if the point responding hit rates and false alarm rates described at is a TCG (NTCG) case depending on whether f . 0(f , the beginning of this section. Doing so, the screening 0). The LDA calculates the coefficient values a1, a2, ..., step results in hit rates of 95%, 95%, and 96%, and false an so that the distances between the mean parameter alarm rates of 24%, 11%, and 25% for the Atlantic, values for groups of TCG cases and NTCG cases are eastern Pacific, and western Pacific basins, respectively. maximized in standard deviation units. Although the hit rates are high by design, the false Not all of the parameters listed in section 3a were alarm rates in conjunction with the large number of used as LDA inputs. Although there is no strict inde- NTCG data points left in the sample set make screening pendence requirement for LDA input parameters, both

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC 462 WEATHER AND FORECASTING VOLUME 24

TABLE 3. The 2 3 2 contingency tables for an idealized forecast from Wilks (2006) (first column) and for the Atlantic basin (second column), eastern Pacific basin (third column), and western Pacific basin (fourth column) after the screening step (first two rows) and after the combined screening step and LDA step (last two rows).

Wilks (2006) Atlantic Eastern Pacific Western Pacific Observed Observed Observed Observed Yes No Yes No Yes No Yes No Screening step Forecast Yes ab665 428 553 479 174 660 476 183 188 No cd36 1 390 430 26 1 462 263 22 553 636 Screening 1 LDA step Forecast Yes ab99 1673 115 1326 43 297 No cd602 1 817 310 390 1 635 597 455 736 527

LAT and CSST were omitted because their contribu- holds especially true for the eastern Pacific basin, which tions were already captured in the climatological for- has the highest frequency of TCG per unit area of any mation probability parameter, CPROB. In addition, region worldwide (Elsberry et al. 1987). In the western results from an initial LDA showed that THDEV and Pacific, CPROB is the third largest contributor, sug- BTWARM have relatively weak contributions to the gesting that the timing and location of TC formation discriminant function. This agrees with Molinari et al. regions may vary more within a given season than in (2000), who suggested that the thermodynamic condi- the other two basins. This may be due to the large intra- tions (in the eastern Pacific basin) are sufficient in and interseasonal variations in the strength and location most places most of the time, and that the best predic- of the monsoon trough in the western Pacific as com- tors of TCG are those associated with an initial distur- pared to the eastern Pacific, which is strongly tied to bance. Hence, the thermodynamic predictors of TC formation in both of those basins (Briegel and Frank BTWARM and THDEV were not used in the final 1997). LDA. This left seven discriminating parameters: The strong contribution by CIRC is consistent with PLAND, DSTRM, CPROB, VSHEAR, CIRC, the idea that an initial disturbance is necessary for TC HDIV, and PCCOLD. formation. Results from two recent studies by Frank Applying LDA yielded a set of seven discriminant and Roundy (2006) and Davis et al. (2007) also support coefficients, whose standardized values (i.e., the pooled this idea. Frank and Roundy (2006) demonstrated a standard deviation for each input parameter times strong relationship between TC formation and several the corresponding coefficient) are shown in Table 4. types of atmospheric waves. They suggested a forcing The standardized discriminant coefficients represent the mechanism whereby waves enhance local circulations, magnitude of each input parameter’s contribution to which in turn promote TC formation. Davis et al. (2007) the discriminant function value. Table 4 shows that the studied vortices in the eastern Pacific and found that parameters contributing the most to the discriminant those that eventually developed into tropical cyclones function values are CPROB, CIRC, and DSTRM. For tended to be stronger and deeper from the outset as the Atlantic and eastern Pacific basins, the highest compared to nondeveloping vortices, once again sug- contribution comes from CPROB, then CIRC, followed gesting a correlation between positive low-level circu- by DSTRM. The large contribution by CPROB suggests lation and subsequent TC development. that these basins each have relatively consistent clima- At first, the finding that DSTRM is a strong contrib- tological patterns of TCG location and timing. This utor to TC formation likelihood was unexpected. This is especially true in light of the finding that VSHEAR had

TABLE 4. Standardized discriminant coefficients for each a much smaller contribution to TC formation proba- parameter in each basin, derived from linear discriminant analysis bility than many parameters, despite the considerable after screening. attention it has received in TC formation research. It is Atlantic Eastern Pacific Western Pacific known that a large amount of vertical shear is detri- mental to TC genesis (Riehl 1948) while a little bit of PLAND 20.21 20.27 20.20 vertical shear may be helpful (McBride and Zehr 1981; CPROB 1.47 1.67 0.88 VSHEAR 20.36 20.22 20.17 Bracken and Bosart 2000). However, it is still not CIRC 1.32 1.41 1.59 known if variations in intermediate values of shear are PCCOLD 0.62 0.29 0.55 important to TC formation like they are to TC intensity HDIV 20.22 20.39 20.33 change. The relatively small contribution of VSHEAR DSTRM 0.75 0.67 1.14 in this analysis suggests that the predictive relationship

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC APRIL 2009 S C H U M A C H E R E T A L . 463 between vertical shear and TC formation is either too 4. Algorithm verification weak or too complex to project significantly on the The algorithm developed in this study provides a LDA. However, it is well known that large amounts of probabilistic, short-term forecast of tropical cyclone vertical shear, such as the shear induced by an existing formation. Since TC formation is an extremely rare TC, have a negative effect on TC formation. Combined event, the typical statistical verification methods that with the effects of storm-induced upwelling and SST would be applied to a probabilistic forecast have biases cooling, the local environment of an existing TC is that make many of them insufficient for estimating skill. generally unfavorable for new TC formation. These For example, consider that for a probability threshold negative impacts are relatively localized about the ex- p 5 0.4% chosen for the Atlantic basin, the corre- isting TC and will decrease with distance, which would sponding hit rate and false alarm rate are HR 5 65.2% explain the relatively large positive values of the stan- and FAR 5 2.3%. These values are not far off from dardized discriminant coefficients for DSTRM. those reported in HH03 and PL86, and at face value LDA produces a binary classification scheme that appear to suggest a skillful forecast. However, with designates each data point as being a TCG or NTCG 428 553 NTCG cases in the Atlantic dataset, this equa- case. Binary classification schemes can be evaluated tes to more than 9800 false alarms. Within the gener- using hit rates and false alarm rates, which are shown for alized framework of this analysis TC genesis is an the combined screening step and LDA step in the last extremely rare occurrence, which makes is difficult to two rows of Table 3. The LDA step leads to a decrease compare existing results that rely on restricted initial in the number of TCG cases correctly identified by at sample sets. As such, we will attempt to define an ap- least a factor of 4. This results in a reduction of hits rates propriate methodology for evaluating the skill of our from values ranging from 94% to 95% to values as low TC formation estimation scheme. as 9%, suggesting that the LDA step greatly reduces the Although standard verification measures for proba- algorithm’s ability to identify TCG occurrences. How- bilistic forecasts are not suitable for use with extremely ever, the number of NTCG cases mistakenly identified rare events, there are some variations of these metrics as TCG cases (i.e., false alarms) by the algorithm that compare the performance of a forecast to that of a dropped even more drastically, reducing false alarm reference forecast to determine skill. The best available rates from 10%–25% to 0.04–0.09%. In the case of this reference forecasts in this case are the null forecast extremely rare occurrence, the large reduction in false (assuming p 5 0% everywhere) and CPROB. For each alarm rates outweighs the lesser reduction in hit rates, verification measure used, it was found that the clima- making LDA an important step in generating a useful tological formation probability forecast performed TC formation prediction scheme. better than the null forecast. As such, climatology was For this project, a probabilistic scheme is sought for used as the reference forecast for each skill score com- estimating TC formation likelihood. This probabilistic puted in the following sections. The results of each scheme was created from the results of the LDA by verification method are given for both the dependent partitioning the TCG cases into 10 subgroups of equal (1995–2005) and independent (2006–07) datasets. size based on their discriminant values, whose values define 10 discriminant value intervals. The TC forma- tion occurrence frequencies were then calculated for a. Brier skill score each interval and used as TC formation probabilities. The first verification diagnostic used was the Brier So, for any given independent data point, the corre- skill score, BSS 5 1 2 (BS/BSref), where BS is the sponding discriminant value is linearly interpolated3 to Brier score of the algorithm and BSref is a reference obtain a formation probability, thus generating a Brier score (Murphy 1973). The Brier score, as defined probabilistic forecast. The next section will evaluate the by Wilks (2006), is performance of this algorithm using statistical forecast verification methods. 2 BS 5 1/n åk51,n ( yk 2 ok) ,

3 The linear interpolation scheme introduced a bias toward where yk and ok are the predicted and observed prob- overprediction in each basin. To correct for this bias, each proba- abilities for case k, respectively, and n is the number of bility is multiplied by the scaling factor So/Sp, where Sp is the forecast–event pairs. The Brier score is essentially the summation of the algorithm-predicted probabilities over all de- mean squared error of the probability forecasts (Wilks pendent data points and So is the summation of all observed TC cases within the dependent dataset. Note that the scaling factor was 2006). The BSS is a measure of the improvement of a calculated and applied independently for each of the three basins. probabilistic forecast over the reference forecast and

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC 464 WEATHER AND FORECASTING VOLUME 24

FIG. 5. ROC plots for the (a) Atlantic, (b) eastern Pacific, and (c) western Pacific basins over the dependent dataset. can range from 2‘ to 1, with 0 indicating no skill with rithm, due to known problems with using the ROC for respect to the reference forecast and 1 being a perfect extremely rare events (Stephenson 2004). Hence, for score. When the climatological formation probability is verification purposes we will look at the ROC skill score used for BSref, the BSS for the dependent years are (SSROC), which is defined in Wilks (2006) as SSROC 5 0.012, 0.024, and 0.025 for the Atlantic, eastern Pacific, (ROCalg 2 ROCref)/(ROCperf 2 ROCref), where ROCalg and western Pacific basins, respectively. The corre- is the ROC value of the product algorithm, ROCref is sponding independent years’ BSSs are 0.006, 0.023, and the ROC value of the reference forecast (in this case

0.017, respectively. Although the BSS values are rela- climatology), and ROCperf is the ROC value corre- tively small, they are all positive, which indicates that sponding to a perfect forecast, which in this case is equal the algorithm provides a more skillful forecast than to 1. Like the Brier skill score, a perfect score is 1 with climatology alone. an SSROC . 0 indicating the product forecast has skill with respect to the reference forecast. The dependent b. Relative operating characteristic skill score years’ SSROC values for the Atlantic, eastern Pacific, Another gauge of forecast skill that can be used for and western Pacific basins are 0.40, 0.08, and 0.58, re- this scheme is the relative operating characteristic spectively. The corresponding SSROC values for the in- (ROC). A ROC diagram is created by plotting the hit dependent years are 0.31, 0.45, and 0.33, respectively. rate versus the false alarm rate using a set of increasing Like the Brier skill score results, all of the SSROC values probability thresholds to make the yes–no decision are positive, indicating the algorithm possesses skill (Mason and Graham 1999; Wilks 2006). The area under over climatology alone. the resultant curve is the ROC value and provides in- c. Reliability diagrams formation regarding the ability of a forecast to dis- criminate between events and nonevents. ROC values To determine how well the algorithm-estimated TC can range from 0 to 1, where a score of 0.5 or less in- formation probabilities correspond to observed frequen- dicates no skill and a score of 1 is perfect. The ROC cies, the reliability diagrams for both the dependent and values for dependent years for the Atlantic, eastern independent years were plotted. Reliability diagrams for Pacific, and western Pacific basins are 0.80, 0.88, and both dependent and independent years for each basin 0.89, respectively (Fig. 5). As discussed at the beginning are shown in Fig. 6. Each diagram includes a distribu- of this section, these scores alone are not necessarily tion plot showing the number of cases belonging to each representative of the overall forecast skill of the algo- probability bin. Examination of the diagrams for the

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC APRIL 2009 S C H U M A C H E R E T A L . 465

FIG. 6. Reliability diagrams for the (a) Atlantic, (b) eastern Pacific, and (c) western Pacific basins. Each plot shows reliability diagrams for (left) dependent years and (right) independent years. dependent years shows that all three basins are generally Formation Probability (TCFP) product, which is cur- well calibrated. There appears to be a slight conditional rently running in real time online (http://www.ssd.noaa. bias in the eastern Pacific toward underprediction of TC gov/PS/TROP/genesis.html). The TCFP product Web formation at higher predicted probabilities, and there site includes a domain-wide overview plot indicating re- may be hint of the same bias in both the Atlantic and gions of elevated TC formation probability along with a western Pacific as well. The diagrams for the independent corresponding enhanced water vapor loop. Each of the years indicate there was a bias toward overprediction in three basins has its own Web page that displays color the Atlantic, a conditional bias toward underprediction contour x–y plots of real-time, climatological, and anom- in the eastern Pacific at higher probabilities, and good aly formation probability and predictor values (Fig. 7). calibration and resolution in the western Pacific. Each basin has been divided into a set of subbasins as shown in Fig. 8. To provide continuity over time, both the product-estimated and climatological formation 5. The NESDIS Tropical Cyclone Formation probabilities (input parameter values) are summed Probability product (averaged) over each subbasin and displayed as time The algorithm developed in this study has been imple- series plots. These time series are useful for identifying mented as the National Environmental Satellite, Data, subbasin-scale variations in input parameter values and and Information Services (NESDIS) Tropical Cyclone the corresponding changes in TC formation probability.

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC 466 WEATHER AND FORECASTING VOLUME 24

FIG. 7. Example of the TCFP product output during the formation of TS Wukong at 1200 UTC 12 Aug 2006. This includes (top left) the main Web page plot denoting areas of enhanced TC formation probability, the predicted TC formation probabilities for the western Pacific basin at (top right) 12 h prior to and (bottom right) 12 h after formation, and (bottom left) the time series plot of cumulative TC formation probability for subbasin WP2. Note that the arrow in the time series plot is pointing to the probability peak coincident with the formation of TS Wukong.

The time series plot of 24-h-predicted TC formation 1998 to 2005, 81.5% of the TCG points occur at times probability over the eastern Pacific (Fig. 8) subbasin is when the subbasin formation probability is greater than shown in Fig. 9. The red open circles in Fig. 9 represent its climatological value and over half of the TCG points the times within 24 h prior to observed TC formation. are coincident with a local maximum in the TC forma- As Fig. 9 shows, most of the red open circles are coin- tion probability. It turns out that this result is robust cident with peaks in subbasin summation TC formation over most subbasins, with 80.3% of the total TCG probability. In fact, for the eastern Pacific subbasin from points occurring at times when the subbasin formation

FIG. 8. The 12 TCFP subbasins.

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC APRIL 2009 S C H U M A C H E R E T A L . 467

FIG. 10. Plot of expected vs observed number of TC formations for each year in the analysis dataset.

FIG. 9. Time series plot of 24-h product predicted TC formation probability (blue line), 24-h climatological TC formation proba- region in the Atlantic, eastern Pacific, and western Pa- bility (black line), and times when TC formation occurred within cific tropical basins was developed. This product uses 24 h (red open circles) for the eastern Pacific subbasin in 2006. both environmental parameters from ATCF best tracks and NCEP GFS data fields as well as convective pa- probability is greater than its climatological value. The rameters from the GOES-E, GOES-W, and MTSAT-1R fine resolution of the analysis grid introduces the prob- water vapor imagery. This algorithm was developed lem of TCs moving in and out of grid boxes during using a two-step process that involves 1) screening out the forecast period, which leads to a negative bias in data points where TC formation is highly unlikely and algorithm hit rates. Using the subbasin summation of 2) linear discriminant analysis. The resultant discrimi- formation probability reduces this problem and gives a nant function values are interpolated to 24-h TC for- more representative view of formation probability when mation probabilities. a region of increased TC formation probability spans a Verification of the TCFP product, using both the large region, such as in an active phase of the monsoon Brier skill score and relative operating characteristic, trough or an elongated easterly wave. showed that the product forecast possesses skill with After examining subbasin averages and sums, it seems respect to both the null hypothesis and climatology. prudent to examine these probabilities on the basin Reliability diagrams also show the product estimated scale. To do so, the product probabilities were summed probabilities to have generally good calibration, with over the entirety of each basin for each dependent and some bias toward underprediction in the eastern Pacific independent year and compared to the observed num- basin. Although individual product probability values ber of TC formations in that basin. The results from this are generally no greater than 10%–15%, they are a vast computation are plotted in Fig. 10. Although the improvement on climatological values, which are on the probabilities have been bias corrected so that the cu- order of 0.1%. In addition, the mean algorithm-derived mulative predicted probabilities equal the total number 24-h TC formation probabilities for TCG cases are of TC observations over the dependent dataset, the 1.5%, 3.0%, and 2.6% for the Atlantic, eastern Pacific, difference between the predicted and actual number of and western Pacific basins, respectively, while the cor- TCs for individual years is apparent in Fig. 10. Most of responding means for the NTCG cases are 0.04%, the data points in Fig. 10 are relatively close to the x 5 y 0.03%, and 0.07%. The differences between the group line (dashed), indicating that the TCFP product is per- means were found to be significant at the 99% confi- forming well on the annual, basin-wide scale. dence level using the Student’s t statistic. b. Future work 6. Summary and future work Several different approaches for improving the TCFP a. Summary product are currently being considered. The first in- A product for estimating the 24-h probability of TC volves extending the analysis domain to include the In- formation within each 58358 latitude–longitude sub- dian Ocean and Southern Hemisphere tropical cyclone

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC 468 WEATHER AND FORECASTING VOLUME 24 basins, essentially making the product global. The real- Tropical cyclone formation is rare at low latitudes (Fig. time, objective TC formation guidance supplied by this 2), where the Coriolis parameter (and hence planetary product could be of use to meteorological agencies in vorticity) is nearly zero and surface winds (and hence these regions, including the JTWC; National Weather pressure gradients) are generally weak (Gray 1975, Service (NWS) offices in Pago Pago and American Samoa; 1979). For this study, the latitude (LAT) is defined at the Fiji TC Regional Meteorological Specialized Center the center of each 58358 latitude–longitude subregion. (RMSC); the La Reunion RMSC; and the Perth, Darwin, and Brisbane TC Warning Centres. b. Percent land Another improvement involves identifying additional Energy fluxes between a tropical cyclone and the environmental parameters to add to the algorithm. Two ocean surface play a crucial role in TC formation and such parameters being considered are the positions intensification (Malkus and Riehl 1960; Leipper 1967; and Dvorak intensities for invest tropical systems sup- Gray 1975). For this reason, it is highly unlikely for a plied by the NCEP/TPC Tropical Analysis and Forecast tropical cyclone to form over land (Fig. 2). For these Branch. Additionally, the current TCFP product de- reasons, the percent of each 58358 latitude–longitude termines the probability of TC formation within 24 h, subregion over land (PLAND) was included as a pa- which is a short forecast period. Future research will rameter. investigate ways in which the forecast period can be extended to 48 h and beyond. This may include the use c. Distance to an existing tropical cyclone of the GFS forecast fields in addition to the analyses. An existing TC introduces storm-related vertical Also, Frank and Roundy (2006) demonstrated that TC shear, changes the local atmospheric moisture content, formation is strongly related to several types of atmo- and causes upwelling that decreases SST, all of which spheric waves and identified corresponding phase rela- result in an environment that is unfavorable for new TC tionships. Given the fact that convective anomalies development. Analysis of the 1949–2005 best tracks associated with these waves are detectable up to a confirms that in the last 57 yr, no TC has ever formed month in advance of TC formation, the statistical TC within 400 km of an existing TC, justifying the inclusion formation forecasts developed by Frank and Roundy of this distance as a parameter. The distance from the (2006) introduce the idea of using upstream convective center of each 58358 latitude–longitude subregion to parameters to extend the TCFP product forecast period. an existing TC (DSTRM) is calculated from the ATCF Work is currently under way to collect and analyze best-track positions for each analysis time. Since the global geostationary satellite imagery for potential up- effects of an existing TC are localized, variations in stream wave–related convective signatures that can be distances beyond 1000 km are not expected to have any incorporated into the current algorithm. predictive quality and hence all values of DSTRM greater than 1000 km are set equal to 1000 km. Acknowledgments. This work is supported by NOAA Grant NA17RJ1228. The authors thank Lixion Avila, d. Sea surface temperature Ed Rappaport, Chris Landsea, and Edward Fukada for providing motivation and guidance in developing this Shapiro and Goldenberg (1998) demonstrated that product; Bernadette Connell for supplying GOES-West SSTs have a direct effect on enhancing TC formation. water vapor imagery; and Ray Zehr for supplying GMS- The exact threshold value for SSTs that support TCG is 05 and MTSAT-1R water vapor imagery from the still unknown; however, Palme´n (1948) suggested an tropical RAMSDIS archives. The views, opinions, and SST of 268C was needed to support tropical cyclone findings in this report are those of the author and should formation. SST, however, is expected to be highly cor- not be construed as an official NOAA or U.S. govern- related with the equivalent potential temperature at 850 ment position, policy, or decision. hPa, which is a large contributor to the THEDEV pa- rameter (see section g of this appendix). As such, the monthly Levitus climatological SST (Levitus 1982) values APPENDIX were used in this algorithm and were only employed for screening purposes. CSST was derived by linearly inter- Environmental and Convective Parameters preting between these monthly values. a. Latitude e. Vertical shear Latitude can have an important impact on TC for- In a study of western Pacific typhoons, Riehl (1948) mation, particularly in terms of limiting its likelihood. found that TC formation can be inhibited by strong

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC APRIL 2009 S C H U M A C H E R E T A L . 469 vertical shear of the horizontal wind. In addition, sev- 200 hPa eral studies have suggested that some weak amount of à THDEV 5 å wt(p)[uE(850 hPa) À uE(p)], vertical shear is necessary for TC development (e.g., p5850 hPa à McBride and Zehr 1981; Bracken and Bosart 2000). where wt(p) is the pressure weighting, and uE and uE This relationship between vertical shear and TC for- are the equivalent potential temperature and saturated mation was recognized by Gray (1968) and DeMaria equivalent potential temperature, respectively, aver- et al. (2001), and was included as dynamical parameters aged over a 88388 subgrid of the GFS data field. in their respective TC genesis parameters. For this study, the magnitude of the vertical shear of the h. Low-level divergence zonal wind (VSHEAR) is calculated from the GFS Several studies of TC formation climatology suggest analysis horizontal wind fields for each subregion as that TC formation is preferred in regions of large-scale qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 2 enhanced low-level convergence. In particular, TC VSHEAR 5 (u850 À u200) 1(y850 À y200) , formation is often associated with enhanced conver- gence zones associated with equatorial Rossby waves y y where u850 ( 850) and u200 ( 200) are the zonal (meridi- (Molinari et al. 2007), the Madden–Julian oscillation onal) winds at 850 hPa and 200 hPa, respectively, aver- (Maloney and Hartmann 2001; Molinari and Vollaro 83 8 aged over the 12 12 latitude–longitude GFS subgrid 2000), and the monsoon trough (Briegel and Frank that is centered on the analysis subregion. 1997). Based on these findings, the 850-hPa horizontal divergence was included as a parameter. The 850-hPa f. Low-level circulation divergence (HDIV) for each subregion in the algorithm Gray (1968, 1975, 1979) noted tropical cyclones domain is calculated from the GFS wind fields: tend to form in regions of anomalous positive vortic- du dv ity. The observational studies of Frank and Clark HDIV 5 1 , (1980), McBride and Zehr (1981), and others confirm dy dx the importance of low-level cyclonic circulation and vorticity in determining formation potential. This in- where du/dy and dy/dy are the values du/dy and dy/dy, crease enhances the heat and moisture fluxes from the respectively, averaged over a 1283128 subgrid of the ocean surface and, in the presence of deep convective GFS data field that is centered on the algorithm sub- bursts, leads to the development of the low-level, warm- region. core TC vortex (Hendricks et al. 2004; Montgomery et al. 2006; Hendricks and Montgomery 2006). The 850- i. Cloud-cleared water vapor brightness temperature hPa circulation (CIRC) is calculated as the line integral of the 850-hPa GFS wind field over the perimeter of an Midlevel moisture levels may influence the develop- 88388 subgrid of the GFS data field that is centered on ment of tropical cyclones by modulating convective the algorithm subregion, so activity. Dry air at middle levels can have a negative I impact on convective activity by reducing the updraft buoyancy via entrainment and decreasing the precipi- CIRC 5 U Á dl, P tation efficiency within developing systems (Ruprecht and Gray 1974; Gray 1975; Bister and Emanuel 1997). where U is the average magnitude of the along-contour As described by Moody et al. (1999), the geostationary component of the horizontal wind field and P is the satellite water vapor imagery in cloud-free regions is contour defined along the perimeter of the 88388 sensitive to the relative humidity in the mid- to upper subgrid. troposphere. Hence, the average brightness temp- erature (BTWARM) within each 58358 latitude– g. Vertical instability longitude subregion is calculated after removing all pixels colder than 2408C and is used as an approxi- Vertical instability is needed to support the deep mation of the midlevel moisture. convective activity necessary for TC formation. The ver- tical instability parameter from DeMaria et al. (2001), j. Percent cold pixel coverage which was developed using the parcel calculation for- mulation described in Ooyama (1990), is used in this Observational studies have found that deep convective study. Vertical instability (THDEV) is calculated for each bursts are often a precursor to TC formation (Gentry subregion using the equation et al. 1970; McBride and Zehr 1981; Zehr 1992; Molinari

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC 470 WEATHER AND FORECASTING VOLUME 24 et al. 2004). To include this environmental signature in Gray, W. M., 1968: Global view of the origin of tropical distur- our algorithm, a parameter to quantify deep convective bances and storms. Mon. Wea. Rev., 96, 669–700. activity was developed. This parameter (PCCOLD) uses ——, 1975: Tropical cyclone genesis in the western North Pacific. Tech. Paper 16-75, Dept. of Atmospheric Sciences, Colorado water vapor (6.7 mm) brightness temperatures and deter- State University, 66 pp. [Available from Dept. of Atmo- mines the percent of pixels within each 58358 latitude– spheric Sciences, Colorado State University, Fort Collins, CO longitude subregion that is colder than 2408C. 80523.] ——, 1979: Hurricanes: Their formation, structure and likely role in k. Climatological 24-h TC formation probability the tropical circulation. Meteorology over the Tropical Oceans, D. B. Shaw, Ed., Royal Meteorological Society, 155–218. Monthly climatological TC formation probabilities ——, 1984: Atlantic seasonal hurricane frequency. Part I: El Nin˜ o were calculated by adding up the total number of TC and 30 mb quasi-biennial oscillation influences. Mon. Wea. formations in each 58358 latitude–longitude subregion Rev., 112, 1649–1668. for each month over the 1949–2005 best-track dataset Harr, P. A., cited 2007: Joint hurricane testbed: Final report— Objective and automated assessment of operational global and then dividing by the number of years (57). These forecast predictions of tropical cyclone formation and life cycle. monthly values were then divided by the total number [Available online at http://www.nhc.noaa.gov/jht/final_rep/ of days in each month to obtain an average daily cli- ENharr_JHTfinalreport_03-05.pdf.] matological TC formation probability for each month. Hendricks, E. A., and M. T. Montgomery, 2006: Rapid scan views Finally, the 24-h climatological probability parameter of convectively generated mesovortices in sheared Tropical Cyclone Gustav (2002). Wea. Forecasting, 21, 1041–1050. (CPROB) was calculated by linearly interpolating be- ——, ——, and C. A. Davis, 2004: The role of ‘‘vortical’’ hot tween these average daily probabilities. towers in the formation of Tropical Cyclone Diana (1984). J. Atmos. Sci., 61, 1209–1232. Hennon, C. C., and J. S. Hobgood, 2003: Forecasting tropical cy- REFERENCES clogenesis over the Atlantic basin using large-scale data. Mon. Beven, J. L., 1999: The boguscane–A serious problem with the Wea. Rev., 131, 2927–2940. NCEP Medium-Range Forecast model in the Tropics. Pre- Kalnay, E., and Coauthors, 1996: The NCEP/NCAR 40-Year prints, 23rd Conf. on Hurricanes and Tropical Meteorology, Reanalysis Project. Bull. Amer. Meteor. Soc., 77, 437–471. Dallas, TX, Amer. Meteor. Soc., 845–848. Knaff, J. A., T. A. Cram, A. B. Schumacher, J. P. Kossin, and M. Bister, M., and K. A. Emanuel, 1997: The genesis of Hurricane DeMaria, 2008: Objective identification of annular hurri- Guillermo: TEXMEX analyses and a modeling study. Mon. canes. Wea. Forecasting, 23, 17–28. Wea. Rev., 125, 2662–2682. Leipper, D. F., 1967: Observed ocean conditions and Hurricane Bracken, W. E., and L. F. Bosart, 2000: The role of synoptic-scale Hilda, 1964. J. Atmos. Sci., 24, 182–186. flow during over the North Atlantic Levitus, S., 1982: Climatological Atlas of the World Ocean. NOAA Ocean. Mon. Wea. Rev., 128, 353–376. Prof. Paper 13, 173 pp. Briegel, L. M., and W. M. Frank, 1997: Large-scale forcing of Malkus, J. S., and H. Riehl, 1960: On the dynamics and energy tropical cyclogenesis in the western North Pacific. Mon. Wea. transformation in steady-state hurricanes. Tellus, 12, 1–20. Rev., 125, 1397–1413. Maloney, E. D., and D. L. Hartmann, 2001: The Madden–Julian Davis, C., C. Snyder, and A. C. Didlake Jr., 2008: A vortex-based oscillation, barotropic dynamics, and North Pacific tropical perspective of eastern Pacific tropical cyclone formation. cyclone formation. Part I: Observations. J. Atmos. Sci., 58, Mon. Wea. Rev., 136, 2461–2477. 2545–2558. DeMaria, J., A. Knaff, and B. H. Connell, 2001: A tropical cyclone Mason, S. J., and N. E. Graham, 1999: Conditional probabilities, genesis parameter for the tropical Atlantic. Wea. Forecasting, relative operating characteristics, and relative operating 16, 219–233. levels. Wea. Forecasting, 14, 713–725. DeMaria, M., M. Mainelli, L. K. Shay, J. A. Knaff, and J. Kaplan, McBride, J. L., 1995: Tropical cyclone formation. Global Perspectives 2005: Further improvements in the Statistical Hurricane In- on Tropical Cyclones, R. L. Elsberry, Ed., WMO/TD 693, Rep. tensity Prediction Scheme (SHIPS). Wea. Forecasting, 20, TCP-38, World Meteorological Organization, 63–105. 531–543. ——, and R. M. Zehr, 1981: Observational analysis of tropical Elsberry, R. L., W. M. Frank, G. J. Holland, J. D. Jarrell, and R. L. cyclone formation. Part II: Comparison of non-developing Southern, 1987: A Global View of Tropical Cyclones. Uni- versus developing systems. J. Atmos. Sci., 38, 1132–1151. versity of Chicago Press, 185 pp. Molinari, J., and D. Vollaro, 2000: Planetary- and synoptic-scale Emanuel, K. A., and D. S. Nolan, 2004: Tropical cyclone acti- influences on eastern Pacific tropical cyclogenesis. Mon. Wea. vity and global climate. Preprints, 26th Conf. on Hurricanes Rev., 128, 3296–3307. and Tropical Meteorology, , FL, Amer. Meteor. Soc., ——, ——, S. Skubis, and M. Dickinson, 2000: Origins and mech- 240–241. anisms of eastern Pacific tropical cyclogenesis: A case study. Frank, N. L., and G. Clark, 1980: Atlantic tropical systems of 1979. Mon. Wea. Rev., 128, 125–139. Mon. Wea. Rev., 108, 966–972. ——, ——, and K. L. Corbosiero, 2004: Tropical cyclone formation Frank, W. M., and P. E. Roundy, 2006: The role of tropical waves in a sheared environment: A case study. J. Atmos. Sci., 61, in tropical cyclogenesis. Mon. Wea. Rev., 134, 2397–2417. 2493–2509. Gentry, R. C., T. T. Fujita, and R. C. Sheets, 1970: Aircraft, ——, K. Lombardo, and D. Vollaro, 2007: Tropical cyclogenesis spacecraft, satellite and radar observations of Hurricane within an equatorial Rossby wave packet. J. Atmos. Sci., 64, Gladys, 1968. J. Appl. Meteor., 9, 837–850. 1301–1317.

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC APRIL 2009 S C H U M A C H E R E T A L . 471

Montgomery,M.T.,M.E.Nicholls,T.A.Cram,andA.B. Sciences, Colorado State University, 91 pp. [Available from Saunders, 2006: A vortical hot tower route to tropical cy- Dept. of Atmospheric Sciences, Colorado State University, clogenesis. J. Atmos. Sci., 63, 355–386. Fort Collins, CO 80523.] Moody, J. L., A. J. Wimmers, and J. C. Davenport, 1999: Remotely Sampson, C. R., and A. J. Schrader, 2000: The Automated Tropical sensed specific humidity: Development of a derived product Cyclone Forecasting System (version 3.2). Bull. Amer. Me- from the GOES Imager channel 3. Geophys. Res. Lett., 26, 59–62. teor. Soc., 81, 1231–1240. Murphy, A. H., 1973: A new vector partition of the probability Shapiro, L. J., and S. B. Goldenberg, 1998: Atlantic sea surface score. J. Appl. Meteor., 12, 595–600. temperatures and tropical cyclone formation. J. Climate, 11, Ooyama, K. V., 1990: A thermodynamic foundation for modeling 578–590. the moist atmosphere. J. Atmos. Sci., 47, 2580–2593. Simpson, J., J. B. Halverson, B. S. Ferrier, W. A. Petersen, R. H. Palme´n, E. H., 1948: On the formation and structure of tropical Simpson, R. Blakeslee, and S. L. Durden, 1998: On the role of cyclones. Geophysics, 3, 26–38. ‘‘hot towers’’ in tropical cyclone formation. Meteor. Atmos. Pasch, R. J., P. A. Harr, L. A. Avila, J.-G. Jiing, and G. Eliott, Phys., 67, 15–35. 2006: An evaluation and comparison of predictions of tropical Stephenson, D., 2004: Skill measures for forecasts of extreme cyclone cyclogenesis by three global forecast models. Pre- rare events. Abstracts, Climate Extremes and Risk Reduction prints, 27th Conf. on Hurricanes and Tropical Meteorology, Conference, London, United Kingdom, Lighthill Institute Monterey, CA, Amer. Meteor. Soc., 14B.5. [Available online of Mathematical Science, London, England. [Available on- at http://ams.confex.com/ams/pdfpapers/108725.pdf.] line at http://www.ucl.ac.uk/lims/events/LIMSConf30Nov04. Perrone, T. J., and P. R. Lowe, 1986: A statistically derived pre- htm.] diction procedure for tropical storm formation. Mon. Wea. Wilks, D. S., 2006: Statistical Methods in the Atmospheric Sciences. Rev., 114, 165–177. 2nd ed. Academic Press, 467 pp. Riehl, H., 1948: On the formation of typhoons. J. Meteor., 5, 247– Zehr, R. M., 1992: Tropical cyclogenesis in the western north Pa- 264. cific. NOAA Tech. Rep. NESDIS 61, 181 pp. [Available from Ruprecht, E., and W. M. Gray, 1974: Analysis of satellite-observed NOAA/National Environmental Satellite, Data, and Infor- tropical cloud clusters. Tech. Paper 219, Dept. of Atmospheric mation Service, Washington, DC 20233.]

Unauthenticated | Downloaded 10/06/21 01:21 PM UTC