• Nenhum resultado encontrado

Analysis of the Most Intense Explosive Cyclone that Occurred Between 2010 and 2020 in the South Atlantic

N/A
N/A
Protected

Academic year: 2023

Share "Analysis of the Most Intense Explosive Cyclone that Occurred Between 2010 and 2020 in the South Atlantic"

Copied!
15
0
0

Texto

(1)

2781

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

Analysis of the Most Intense Explosive Cyclone that Occurred Between 2010 and 2020 in the South Atlantic

Hugo Nunes Andrade1, André Becker Nunes2 e Rita de Cássia Marques Alves3

1Mestrando em Sensoriamento Remoto no Programa de Pós-Graduação em Sensoriamento Remoto, Universidade Federal do Rio Grande do Sul, Av.

Bento Gonçalves 9500, Campus do Vale, Porto Alegre-RS, CEP: 91501-970 hugonandrade@hotmail.com; 2Dr. em Meteorologia, Professor do Departamento de Meteorologia da Universidade Federal de Pelotas, Campus Universitário, S/N, Capão do Leão, CEP: 96010-610, Pelotas-RS, beckernunes@gmail.com; 3Dra. em Meteorologia, Professora do Programa de Pós-Graduação em Sensoriamento Remoto, Universidade Federal do Rio Grande do Sul, Av. Bento Gonçalves 9500, Campus do Vale, Porto Alegre-RS, CEP: 91501-970 ritacma@terra.com.br.

Artigo recebido em 27/07/2022 e aceito em 07/10/2022 A B S T R A C T

The coastal region of South America is known for the high frequency of extratropical cyclones. From 2010 to 2020, there was an exceptional case regarding intensity, reaching 2.73 Bergeron, between January 02 and 03, 2019. To better understand the characteristics of this type of explosive cyclone, this work sought to investigate the synoptic conditions in this phenomenon. To this end, a visual inspection method of the sea level pressure charts was applied, allied with the functions available in the Grid Analysis and Display software. The cyclone began by transitioning from the continental low to the extratropical cyclone, associated with a trough at higher levels in a zone of weak temperature advection. As the system developed, there was a fracture in the upper air trough, acquiring negative horizontal inclination and the transition of the cyclone from the tropical to the polar side of the jet streak. Sea heat fluxes become relevant only 6 hours after cyclogenesis and enhance as the surface wind increases in the cold sector of the cyclone. In addition, a robust stratospheric ozone intrusion arose close to 700 hPa in the cyclone region, related to the dynamic tropopause folding.

Keywords: Bombogenesis, Heat Fluxes, Dynamic Tropopause, ERA5, Synoptic Analysis.

Análise do ciclone explosivo mais intenso ocorrido no período entre 2010 e 2020 no Atlântico Sul

R E S U M O

A região da costa da América do Sul é conhecida pela alta frequência de ciclones extratropicais. Durante o período de 2010 a 2020, houve um caso excepcional com relação à intensidade, atingindo 2,73 Bergeron, entre 02 e 03 de janeiro de 2019. Para entender melhor as características associadas com este tipo de ciclone explosivo, este trabalho buscou investigar as condições sinóticas ocorridas neste fenômeno. Para tal, foi aplicado um método de inspeção visual das cartas de pressão ao nível médio do mar, aliado a funções disponíveis no software Grid Analysis and Display. O ciclone iniciou através da transição da baixa continental ao ciclone extratropical associado a um cavado em níveis superiores em uma zona de fraca advecção de temperatura. Conforme o sistema se desenvolveu, houve uma fratura no cavado em ar superior, adquirindo inclinação horizontal negativa e a transição do ciclone do lado tropical para o polar do jet streak. Os fluxos de calor do oceano se tornaram relevantes apenas 6 horas após a ciclogênese e se destacam com o aumento do vento à superfície no setor frio do ciclone. Além disso, observou-se uma forte intrusão de ozônio estratosférico até próximo de 700 hPa na região do ciclone, relacionado à dobra da tropopausa dinâmica.

Palavras-chave: Bombogênese, Fluxos de Calor, Tropopausa Dinâmica, ERA5, Análise Sinótica.

Introduction

Cyclones are a crucial component in the dynamics of the atmosphere. They develop where the flow favors their growth and play a fundamental role in the hydrological cycle since they carry heat, humidity, and momentum (Peixoto and Oort, 1992; Wallace and Hobbs, 2006; Reboita

et al., 2010; Gertler and O’Gorman, 2019;

Gramcianinov et al., 2019; Reale et al., 2019;

Aragão and Porcù, 2022).

A portion of these systems has a disruptive capacity for human activities, especially in coastal regions and on shipping routes, leading, in some

(2)

2782

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

cases, to infringe on life losses. These characteristics are associated with intense winds, large precipitation amounts, dangerous oceanic conditions, and sudden temperature changes (Allen et al., 2010; Reboita et al., 2010; Liberato et al., 2011; Reale and Lionello, 2013; Reale et al., 2019;

Kodama et al., 2019; Gramcianinov et al., 2020).

Explosive cyclones or “bombs” are preferably maritime and cold season phenomena (Sanders and Gyakum, 1980; Bluestein, 1993; Lim and Simmonds, 2003; Bitencourt et al., 2013).

Wang and Rogers (2001) compiled the mechanisms associated with explosive cyclogenesis: intense high levels forcing (cyclonic vorticity advection); stratospheric air intrusion with high potential vorticity; latent heat release;

fluxes from the ocean to the atmosphere;

highlighted local baroclinicity from differential diabatic heating (Uccellini et al., 1985; Rogers e Bosart, 1986; Atlas, 1987; Davis e Emanuel, 1988;

Hirschberg e Fritsch, 1991; Kuwano-Yoshida e Asuma, 2004; 2008; Piva et al., 2011; Kuwano- Yoshida e Enomoto, 2013; Liberato, 2014; Heo et al., 2015; Reboita et al., 2021). Researchers suggest that they are similar to ordinary cyclones, essentially governed by baroclinic instability, but the role of these attributes mentioned above strongly contributes to the rapid deepening. It is important to emphasize that the participation of these properties is dependent on intensity and region, in addition to the stage of development (Wash et al., 1992; Zhang et al., 2017).

Furthermore, the importance of each factor can vary substantially (Reed et al., 1993; Wang and Rogers, 2001; Rudeva and Gulev, 2011).

According to Sanders and Gyakum (1980), an explosive cyclone must meet a simple criterion:

the resulting critical rate exceeds 1 Bergeron, considering the variation of the central pressure in 24 hours and the average latitude in the same period concerning 60ºS (Eq. 1). Some cases may reach high values, in Bergeron, and be categorized as strong systems.

Among these, two famous ones stand out:

the Queen Elizabeth II storm in 1978 and the Presidents’ Day cyclone in 1979. The first was a storm that occurred above 40ºN in the western Atlantic Ocean, characterized by a variation of 60 hPa in 24 hours, clearly underestimated at the time by the meteorological models, in which the cyclone took tropical cyclone characteristics in the pressure, wind and cloud fields. The study divides into two parts, one referring to the initial phase of

low pressure in an area of intense baroclinity and the other about the relationship with the trough at higher levels (Gyakum, 1983a; 1983b). The second refers to a storm that occurred on the east coast of the United States, also in a very baroclinic area, in which latent and sensitive heat fluxes played a substantial role. The contributions of the transient trough with the vorticity advection and intense convergence at low levels also helped in the strong convection associated with the cyclone center. The storm development led to records in the accumulated snow (Bosart, 1981; Bosart and Lin, 1984; Uccellini et al., 1985; Uccellini et al., 1985;

Whitaker et al., 1988).

Heo et al. (2015) studied a case on April 3- 4, 2012, in the Sea of Japan. Using several data sources, the authors showed that strong baroclinic instability, temperature advection associated with the cut-off low at high levels, and interaction with potential vorticity anomalies between high and low levels are essential for explosive cyclogenesis.

Furthermore, sensitivity experiments indicated that latent heat release amplifies intensification, extension, and velocity.

For the Southern Hemisphere, Reale et al.

(2019) indicate that the western part, close to South America, is characterized by higher normalized deepening rates (NDR) than those close to Australia. Avila et al. (2016) used reanalysis data from MERRA-2 (Rienecker et al. (2011) to analyze an intense explosive cyclone in January 2014. They demonstrated that baroclinic instability was the predominant factor in the system intensification.

Ocean heat fluxes were insignificant for the development, and dynamic tropopause anomalies were consistent with the literature (Hoskins, 1985;

Santurette and Georgiev, 2005). These characteristics were also found by Avila et al.

(2021), with more emphasis on the intense ones.

From a broader view, Reis et al. (2020) analyzed an explosive cyclone that occurred in September 2016 near the La Plata Basin and linked its eventuality to the anomalous temperature dipole between Southern Brazil and Antarctic Peninsula.

These violent events are becoming more frequent in an imminent climate change scenario, but there are many uncertainties as to what extent these systems will go through (Catto et al., 2019) since a few gaps in explosive cyclones knowledge need to be filled.

Therefore, this work aims to understand the most intense explosive cyclone that occurred between 2010 and 2020 as part of an effort to

(3)

2783

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

recognize the characteristics interconnected with these exceptional cases, which impact the weather during its displacement.

Data and methods

Data from ERA5 reanalysis, with a spatial resolution of 0,25º x 0,25º with 139 vertical levels (Hersbach et al., 2020), were used. During the first five days of January 2019, data were selected at 3- hour intervals, covering the case under study (02- 03 January). ERA5’s reanalysis benefits from ten years of improvements over the previous one;

ERA-Interim (Dee et al., 2011), mainly in the physics model, dynamics, and data assimilation, in addition to the more refined spatial resolution (Hersbach et al., 2020). Recent studies with different objectives and applied to different regions have utilized this newer reanalysis, for example, Gramcianinov et al. (2020), Aragão and Porcù (2022), and Pezzi et al. (2022). Satellite data from the Geostationary Environmental Satellite (GOES- 16) were also used in the infrared (channel 13 – 10.36 μm) and water vapor (channels 7, 8, and 9 – 6.2 μm; 6.9 μm; 7.3 μm for high, medium and low levels, respectively).

Regarding the cases’ detection, the regions with a surface trough and a closed isobar in the sea level pressure (slp) were first visually identified within the spatial domain (15ºS and 60ºS, 75ºW, and 10ºW).

Then, at the regions previously identified, through the software Grid Analysis and Display (GrADS), a grid was determined that includes them. This grid was scanned line by line from south to north and west to east, rigorously in this order by amin function, available in the software. The minimum value of slp that resulted from this process is given in an X-Y field, and then a conversion to lat/lon is made in the same grid (aminlocy/aminlocx functions).

After tracking the cyclone, from its cyclogenesis to the most intense phase, the NDR, defined by Sanders and Gyakum (1980), is calculated:

𝑁𝐷𝑅 = ∆𝑃

24∗𝑠𝑒𝑛 60º 𝑠𝑒𝑛 𝜑

Where ΔP is the variation of the central pressure in 24 hours and φ is the mean latitude, considering the start and end points of the system’s explosive phase. When this rate is equal to or exceeds 1 Bergeron, the cyclone is a “bomb” or explosive.

The algorithm utilized here is similar to the one used by Zhang et al. (2017), which also identified cyclones through visual inspection in a study for the North Pacific. The authors defined a lat/lon box of 10º x 16º for a 6 hours time resolution. Since the timestep was selected for 3 hours, the slp minimum does not exceed 5º of latitude and longitude concerning the previous timestep to prevent the identification of another system and track the same cyclone. Once the reanalysis employed here has higher temporal and spatial resolutions, the box criterion is justified.

Explosive cyclones were differentiated by intensity: 1,00 ≤ NDR < 1,30 are weak; 1,30 ≤ NDR ≤ 1,80 are moderate, and NDR > 1,80 are intense. This classification differs minimally from the one proposed by Sanders (1986), which discriminates weak explosive cyclones as 1,0 ≤ TNA ≤ 1,2 and moderates as 1,3 ≤ TNA ≤ 1,8. As this work uses two decimal places, weak and moderate elements are minimally altered (Andrade, 2019).

Results and discussion

The following analysis highlights the explosive cyclone that occurred in the summer, between January 2 and 3, 2019. This case stands out because it is the strongest from 2010 to 2020 in the South Atlantic. According to the equation of Sanders and Gyakum (1980), it is characterized as intense, in which the central pressure normalized deepening rate reaches above 1,80 B. The slp variation of the cyclone center acquired 52,09 hPa in 24 hours, resulting in the NDR = 2,73 B. This case would fit into a new categorization proposed by Zhang et al. (2017), where cyclones with rates equal to or above 2,3 Bergeron are called super explosive cyclones.

The first slp minimum detected occurred at 21Z on January 02, where an inverted trough at the surface (Figure 1) was associated with the Northwestern Argentinean Low (NAL) (Seluchi et al., 2003). This system has a thermal origin and is forced orographically by the subsidence of air leeward of the Andes. During the summer, NAL is an almost permanent system and is often associated with transient phenomena due to its location coincident with high-level jets (Seluchi and Saulo, 2012). This transition from the continent trough to the cyclogenesis on the coast or in the ocean is frequent in this region (Caballero et al., 2018).

(4)

2784

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

Figure 1 – Streamlines and wind magnitude (shaded) at 250 hPa and slp (contours in pink) for: a) 21Z on day 2; b) 03Z on day 03; c) 09Z on day 03; d) 15Z on day 03; e)21Z on day 03.

The inverted surface trough, as mentioned earlier, was located downstream of the medium levels difluent flow, indicating that this leeward flow acted as the tendency for surface pressure drop (Sanders, 1993). Pettersen and Smebye (1971)

named this configuration Type B cyclone. In addition, figure 1 shows that, at high levels, there is an amplified trough coming from high latitudes associated with the surface cyclone. There is also a pronounced ridge northward of the subtropical jet.

(5)

2785

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

This configuration points to the encounter of two air masses from different regions, shown in the 500-1000 hPa thickness field (Figure 5) and, more leniently, in the thermal advection field (not shown here) along the surface trough. The extension of the low-pressure area of the continent occurred in colder waters, between 15ºC and 21ºC sea surface temperature, southward of the cyclogenetic region of the Brazil-Malvinas Confluence (not shown here). These colder waters may explain the irrelevant role of surface latent and sensitive heat fluxes during the precyclogenesis and cyclogenesis phases. In this case, the cyclone already acquired the denomination of explosive in this time step, according to the slp variation in 24 hours.

At high levels, the cyclone begins under the region of maximum wind in an amplified trough with positive horizontal inclination. During the development of the system, there is a fracture in such trough, so it acquires a smaller amplitude and negative horizontal tilt. The latter configuration is associated with more significant convective activity and lower static stability (MacDonald, 1976; Glickman et al., 1977;

Bluestein, 1993; Cossetin et al., 2016; Schemm et al., 2020).

The difluence pattern at this level is observed only 9 hours after the cyclone detection and continues until the end of the explosive phase.

The location of this pattern takes place mostly southeastward of the cyclone on the surface in the early stages. It performs the transition northeastward in the final moments of the explosive phase with the split of the subtropical and polar jets (Figure 1d; 1e). Andrade (2019) shows, through composite fields, that the difluence at high levels in intense explosive cyclones has greater emphasis than non-intense (weak and moderate) ones. Thus, the mass convergence at low levels is emphasized, providing deep convection, and intensifying the cyclone.

As the cyclone closes its circulation in the next time step, it moves southeast, according to the maximum positive thermal advection (Bluestein, 1993). Vigorous vertical movement establishes around maximum positive thermal advection, increasing convective activity. Since the beginning, the negative (cold) advection is more intense in magnitude than the positive (warm), indicating more potential for instabilities in the transition region of the cold front through a higher gradient.

In explosive cyclones, the troughs deepen the tropopause over the cold side due to their interaction with air masses. In this case, it is considered that there is a fold in the dynamic tropopause (Uccellini et al., 1985; Avila et al., 2016, 2021; Nunes and Avila, 2017; Heo et al., 2019), which entails a maximum of potential vorticity (PV). For the Southern Hemisphere, cyclonic PV is negative, and according to Santurette and Georgiev (2005), 1,5 units of potential vorticity represents dynamic tropopause.

It is observed in the PV fields at 300 hPa (not shown here) that a tongue extends to the center of the cyclone from 09Z on day 03, where this subsidence is shown with a weak descending movement at 500 hPa in the same region. At the next time step, this vertical movement becomes very pronounced, while the omega values reach between 1,5 and 2,0 Pa/s. At the same time, satellite images in the water vapor channel at medium levels also begin to show this subsidence of stratospheric air (not shown here). This subsidence movement is sustained until the end of the explosive phase and may lower tropopause to even lower pressure levels with the maturation of the cyclone. At the end of this section, vertical sections will be shown to illustrate this vertical movement.

In the period of more preeminent slp variation (15Z of day 03), there is a decrease of 10,84 hPa concerning the previous time step (12Z of day 03). At this moment, we note a mid-level cut-off low (Figure 2) while the cyclone begins to lose its vertical tilt, indicating the final stage of baroclinic development as it becomes the equivalent barotropic type.

The next instant, at 18Z on the 03rd day, high values of sensitive and latent heat fluxes are noted in the cold part of the cyclone, in the southwest quadrant. The latent heat flux has its maximum in the next time step (end of the explosive phase), with significant values over most of the cold side of the cyclone. Negative values indicate the transfer from the sea to the air, while the positives represent the opposite (Figures 3 and 4). These variables begin to impact the cyclone, only 6 hours after cyclogenesis, especially in the southern region of the cyclone, following the transition to the southwest and east during the explosive phase. These results agree with those obtained by Rudeva and Gulev (2011), who also observed the most prominent latent and sensitive heat fluxes in the cold part of the cyclone in the North Hemisphere. These results indicate that, in

(6)

2786

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

this case, ocean fluxes were relevant in the maintenance and intensification of the explosive

cyclone, not acting in the environment for its formation

Figure 2 – Geopotential height at 500 hPa (contours in black) and slp (contours in blue) for: a) 21Z on day 2;

b) 03Z on day 03; c) 09Z on day 03; d) 15Z on day 03; e)21Z on day 03.

(7)

2787

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

Figure 3 – Surface latent heat flux (shaded) in W/m² and slp (purple contours) for: a) 21Z on day 2; b) 03Z on day 03; c) 09Z on day 03; d) 15Z on day 03; e)21Z on day 03.

.

(8)

2788

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

Figure 4 – Surface sensible heat flux (shaded) in W/m² and slp (purple contours) for: a) 21Z on day 2; b) 03Z on day 03; c) 09Z on day 03; d) 15Z on day 03; e)21Z on day 03.

At the end of the explosive phase (21Z day 03), it is noted that there is a trapping of warm air in the center of the cyclone, as seen in the thickness field (Figure 5e). This trapping indicates the

seclusion of warm air as the cold front advances over the warm front due to its density. The configuration presented seems to demonstrate a Shapiro-Keyser system (Reboita et al., 2017; 2022;

(9)

2789

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

Avila et al., 2021), where the cold front becomes more meridional and disconnects from the cyclone.

Heat fluxes reach their maximums in the cold part of the cyclone, as well as wind reach approximately 33 m/s (120 km/h) in the same region, while the system is in phase with the cut-off low at higher levels.

The system’s tracking showed that the cyclone moved over the ocean, covering a sea surface temperature variation of approximately 12ºC (not shown here), a wide range, as emphasized by Sanders and Gyakum (1980) for bomb cases.

Figure 5 – 500-1000 hPa thickness (colored contours) and slp (black contours) for: a) 21Z on day 2; b) 03Z on the day 03; c) 09Z on day 03; d) 15Z on day 03; e)21Z on day 03.

(10)

2790

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

Through the latitudinal vertical sections of the cyclone center, the anomalous dynamic tropopause reaches the level pressure of 700 hPa, with very intense PV in much of the layer above (Figure 6). This anomaly occurs lagged to the west of the surface cyclone, on the cold side. In the most active phase of the system, it is possible to observe higher values of incipient PV at low levels, resulting from the rotation of the cyclone on the surface.

The ozone mass mixing ratio follows the tropopause fold (Figure 7). Through tests, the values of 1,5 (kg.kg-1) for this variable are

compatible with the values of 1,5 PV, indicating the intrusion of stratospheric air rich in ozone. In addition, we can observe in both figures an upstream wavy pattern at higher levels, similar to the gravity waves, on the cold side of the cyclone.

The results established here are consistent with Avila et al. (2016, 2021).

This configuration leads to increased heat fluxes from the ocean, observed above, which can expand the latent heat release (i.e., cloud supplement) and warm the layer. Once said this, the frontal structure developed to the warm-occluded stage (Pang et al., 2022).

Figure 6 – Vertical section of the potential vorticity centered at the latitude of the cyclone center for: a) 21Z on day 02: -39.75º; b) 03Z on day 03: -41.75º; c) 09Z on day 03: -43.75º; d) 15Z on day 03: -46.25º; e) 21Z on day 03: -47º.

(11)

2791

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

Figure 7 – Vertical section of the ozone mass mixing ratio centered at the latitude of the cyclone center for:

a) 21Z on day 02: -39.75º; b) 03Z on day 03: -41.75º; c) 09Z on day 03: -43.75º; d) 15Z on day 03: -46.25º;

e) 21Z on day 03: -47º.

There are no significant slp variations in the following timesteps. While the features mentioned above decrease, the system enters the decay phase the next day.

Conclusions

This study analyzed the South Atlantic’s most intense explosive cyclone between 2010 and 2020, which took place on January 2 and 3, 2019,

(12)

2792

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

and presented a decrease of 52,09 hPa in 24 hours, leading to NDR = 2,73 by ERA5 reanalysis data.

Explosive cyclones of this intensity are rare phenomena near South America.

The cyclone began its transition from continental low to cyclone as NAL, with strong instabilities of convective systems reaching Uruguay and southern Brazil through the organization of the frontal system. In addition, there was dynamic support by the trough at higher levels over a weak thermal horizontal gradient. The cyclone starts over the colder ocean due to the Malvinas current, without influence from sea heat fluxes, and moves through a wide range of sea surface temperatures. The effect of this last feature is indirect once the ocean movements are slower than the atmosphere.

During the transition of the surface cyclone from the north of the jet streak to the south, the system experiences a profound drop in central pressure, with vigorous vertical movements and intrusion of stratospheric air with the folding of the tropopause.

Surface wind maximums occurred in the southwest quadrant, reaching approximately 33 m/s in the most intense period, with latent and sensible heat fluxes being more prominent in this region. The configuration at the end of the explosive phase denoted a Shapiro-Keyser model, with the frontal structure showing a warm occlusion.

For a future study, the simulation of this case in numerical models of weather forecast is suggested to verify whether the parametrizations satisfactorily predict the behaviors of this atypical case.

Acknowledgments

The first author would like to thank the Coordination for the Improvement of Higher Education Personnel - Brazil (CAPES) for the master’s scholarship granted.

References

Allen, J.T.; Pezza, A.B.; Black, M.T., 2010.

Explosive cyclogenesis: A global climatology comparing multiple reanalyses. Journal of Climate, 23, 6468-6484.

Andrade, H. N. Análise dos Casos de Ciclones Explosivos no Atlântico Sul em 2014 e 2015 de Acordo com Diferentes Resoluções Espaciais

de Reanálises. 53 f. Trabalho de Conclusão de Curso (Graduação em Meteorologia) – Universidade Federal de Pelotas, Pelotas, 2019.

Avila, V.D.; Nunes, A.B.; Alves, R.C.M., 2016.

Análise de um caso de ciclogênese explosiva ocorrido em 03/01/2014 no sul do Oceano Atlântico. Revista Brasileira de Geografia Física, 9, 1088-1099.

Avila, V.D.; Nunes, A.B.; Alves, R.C.M., 2021.

Comparing explosive cyclogenesis of different intensities occurred in Southern Atlantic. Anais da Academia Brasileira de Ciências, 93, 1-23.

Aragão, L.; Porcù, F., 2022. Cyclonic activity in the Mediterranean region from a high- resolution perspective using ECMWF ERA5 dataset. Climate Dynamics, 58, 1293-1310.

Bitencourt, D.P.; Fuentes, M.V.; Cardoso, C.S., 2013. Climatologia de ciclones explosivos para a área ciclogenética da América do Sul. Revista Brasileira de Meteorologia, 28, 43-56.

Bluestein, H.B. Synoptic-dynamic meteorology in midlatitudes. Volume II: Observations and theory of weather systems. New York: Oxford University Press, 1993. 594 p.

Bosart, L.F., 1981. The Presidents’ Day snowstorm of 18-19 February 1979. A subsynoptic-scale event. Monthly Weather Review, 109, 1542-1566.

Bosart, L.F.; Lin, S.C., 1984. A diagnostic analysis of the Presidents’ Day Storm of February 1979. Monthly Weather Review, 112, 2148-2177.

Caballero, C.B.; Ogassawara, J.F.; Dorneles, V.R.; Nunes, A.B.; 2018. Precipitação pluviométrica em Pelotas/RS: tendência, sistemas sinóticos associados e influência da ODP. Revista Brasileira da Geografia Física, 11, 1429-1441.

Catto, J.L.; Ackerley, D.; Booth, J.F.; Champion, A.J.; Colle, B.A.; Pfahl, S.; Pinto, J.G.;

Quinting, J.F.; Seiler, C., 2019. The Future of Midlatitude Cyclones. Current Climate Change Reports, 5, 407-420.

Chen, S.; Kuo, Y.; Zhang, P.; Bai, Q., 1992. Notes and Correspondence. Climatology of Explosive Cyclones off the East Asian coast.

Monthly Weather Review, 120, 3029-3035.

Cossetin, F.; Nunes, A.B.; Teixeira, M.S., 2016.

Análise do Movimento Vertical sob Duas Diferentes Configurações de Altos Níveis da Troposfera. Ciência e Natura, 38, 484- 490.

(13)

2793

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

Davis, C.A.; Emanuel, K.A., 1988. Observational evidence for the influence of surface heat fluxes on rapid maritime cyclogenesis.

Monthly Weather Review, 116, 2649-2659.

Dee, D.P.; Uppala, S.M.; Simmons, A.J.;

Berrisford, P.; Poli, P.; Kobayashi, S.; Andrae, U.; Balmaseda, M.A.; Balsamo, G.; Bauer, P.;

Bechtold, P.; Beljaars, A.C.M.; van de Berg, L.; Bidlot, J.; Bormann, N.; Delsol, C.;

Dragani, R.; Fuentes, M.; Geer, A.J.;

Haimberger, L.; Healy, S.B.; Hersbach, H.;

Hólm, E.V.; Isaksen, L.; Kållberg, P.; Köhler, M.; Matricardi, M.; McNally, A.P.; Monge- Sanz, B.M.; Morcrette, J.-J.; Park, B.-K.;

Peubey, C.; de Rosnay, P.; Tavolato, C.;

Thépaut, J.-N.; Vitart, F., 2011. The ERA- Interim reanalysis: configuration and performance of the data assimilation system.

Quarterly Journal of the Royal Meteorological Society, 137, 553-597.

Gertler, C.G.; O’Gorman, P.A., 2019. Changing available energy for extratropical cyclones and associated convection in Northern Hemisphere summer. Earth, Atmospheric, and Planetary Sciences, 116, 4105-4110.

Glickman, T.; MacDonald, N.J.; Sanders, F., 1977. New Findings and the Apparent Relationship between Convective Activity and the Shape of 500 mb Troughs. Monthly Weather Review, 105, 1060.

Gramcianinov, C.B.; Campos, R.M.; Camargo, R.; Hodges, K.I.; Guedes Soares, C.; Silva Dias, P.L., 2020. Analysis of Atlantic extratropical storm tracks characteristics in 41 years of ERA5 and CFSR/CFSv2 databases.

Ocean Engineering, 216, 108111.

Gramcianinov, C.B.; Hodges, K.I.; Camargo, R., 2019. The properties and genesis environments of South Atlantic cyclones. Climate Dynamics, 53, 4115-4140.

Gyakum, J. R., 1983a. On the Evolution of the QE II Storm. I: Synoptic Aspects. Monthly Weather Review 111, 1137-1155.

Gyakum, J. R., 1983b. On the Evolution of the QE II Storm. II: Dynamic and Thermodynamic Structure. Monthly Weather Review 111, 1156- 1173.

Heo, K-Y.; Seo, Y-W.; Ha, K-J.; Park, K-S.; Kim, J.; Choi, J-W.; Jun, K.; Jeong, J-Y., 2015.

Development mechanisms of an explosive cyclone over East Sea on 3-4 April 2012.

Dynamics of Atmospheres and Oceans, 70, 30- 46.

Heo, K-Y.; Ha, K-J.; Ha, T., 2019. Explosive Cyclogenesis around the Korean Peninsula in May 2016 from a Potential Vorticity Perspective: Case Study and Numerical Simulations. Atmosphere, 10, 322, 1-19.

Hersbach, H.; Bell, B.; Berrisford, P.; Hirahara, S.;

Horányi, A.; Muñoz-Sabater, J.; Nicolas, J.;

Peubey, C.; Radu, R.; Schepers, D.;

Simmonds, A.; Soci, C.; Abdalla, S.; Abellan, X.; Balsamo, G.; Bechtold, P.; Biavati, G.;

Bidlot, J.; Bonavita, M.; De Chiara, G.;

Dahlgren, P.; Dee, D.; Diamantakis, M.;

Dragani, R.; Flemming, J.; Forbes, R.; Fuentes, M.; Geer, A.; Haimberger, L.; Healy, S.;

Hogan, R.J.; Hólm, E.; Janisková, M.; Keeley, S.; Laloyaux, P.; Lopez, P.; Lupu, C.; Radnoti, G.; De Rosnay, P.; Rozum, I.; Vamborg, F.;

Villaume, S.; Thépaut, J.N., 2020. The ERA5 global reanalysis. Quarterly Journal of the Royal Meteorological Society, 146, 1999- 2049.

Hirschberg, P.A.; Fritsch, J.M., 1991. Tropopause undulations and the development of extratropical cyclones. Part I: Overview and observations from a cyclone event. Monthly Weather Review, 119, 496-517.

Hoskins, B.J.; Mcintyre, M.E.; Robertson, A.W., 1985. On the use and significance of isentropic potential vorticity maps. Quarterly Journal of the Royal Meteorological Society, 111, 877- 946.

Kodama, C.; Stevens, B.; Mauritsen, T.; Seiki, T.;

Satoh, M., 2019. A New Perspective for Future Precipitation Change from Intense Extratropical Cyclones. Geophysical Research Letters, 46, 12435-12444.

Kousky, V.E.; Elias, M. Meteorologia Sinótica:

Parte I. São José dos Campos: INPE, 1982.118 p.

Kuwano-Yoshida, A.; Asuma, Y., 2004.

Structures and environment of explosively developing extratropical cyclones in the northwestern Pacific region, Monthly Weather Review, 132, 1121-1142.

Kuwano-Yoshida, A.; Asuma, Y., 2008.

Numerical study of explosively developing extratropical cyclones in the Northwestern Pacific region. Monthly Weather Review, 136, 712-740.

Kuwano-Yoshida, A.; Enomoto, T., 2013.

Predictability of explosive cyclogenesis over the Northwestern Pacific region using

(14)

2794

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

ensemble Reanalysis. Monthly Weather Review, 141, 3769-3785.

Liberato, M. L. R., 2014. The 19 January 2013 windstorm over the North Atlantic: large-scale dynamics and impacts on Iberia. Weather Climate Extremes 5, 16–28.

Liberato, M. L. R.; Pinto, J. G.; Trigo, I. F.; Trigo, R. M., 2011. Klaus – an exceptional winter storm over northern Iberia and southern France. Weather 66, 330–334.

Lim, E.P.; Simmonds, I., 2002. Explosive cyclone development in the Southern Hemisphere and a comparison with Northern Hemisphere events. Monthly Weather Review, 130, 2188- 2209.

Lionello, P.; Trigo, I.F.; Gil, V.; Liberato, M.L.;

Nissen, K.M.; Pinto, J.G.; Raible, C.C.; Reale, M.; Tanzarella, A.; Trigo, R.M.; Ulbrich, S.;

Ulbrich, U., 2016 Objective climatology of cyclones in the Mediterranean region: a consensus view among methods with different system identification and tracking criteria.

Tellus A: Dynamic Meteorology and Oceanography, v.68, n.1, 2016.

Macdonald, N., 1976. On the Apparent Relationship Between Convective Activity and the Shape of 500 mb Troughs. Monthly Weather Review, 104, 1618-1622.

Nunes, A.B.; Avila, V.D., 2017. Dynamic Tropopause Anomaly: Case Study in South America. Anuário do Instituto de Geociências, 40, 83-92.

Pang, H.; Fu, G.; Lu, C., 2022. Structure of the Potential Vorticity of an Explosive Cyclone over the Eastern Asian Region in Late November 2013. Journal of Ocean University of China, 21, 253-263.

Peixoto, J.P.; Oort, A.H. Physics of Climate. New York: American Institute of Physics, 520 p.

Pettersen, S.; Smebye, S.J., 1971. On the development of extratropical cyclones.

Quarterly Journal of The Royal Meteorological Society 97, 457-482.

Piva, E. D., GAN, M.A., MOSCATI, M.C.L., 2011. The role of latent and sensible heat fluxes in an explosive cyclogenesis over the South American East Coast. Journal of Meteorological Society of Japan, 89, 637-663.

Pezzi, L.P.; Quadro, M.F.L.; Lorenzzetti, J.A.;

Miller, A.J.; Rosa, E.B.; Lima, L.N.; Sutil, U.A., 2022. The effect of Oceanic South Atlantic Convergence Zone episodes on regional SST anomalies: the roles of heat

fluxes and upper-ocean dynamics. Climate

Dynamics. Disponível em:

https://doi.org/10.1007/s00382-022-06195-3 Reale, M.; Liberato, M.L.R.; Lionello, P.; Pinto,

J.G.; SALON, S.; ULBRICH, S., 2019. A global climatology of explosive cyclones using a multi-tracking approach. Tellus A: Dynamic Meteorology and Oceanography, 71, 1-19.

Reale, M.; Lionello, P., 2013. Synoptic climatology of winter intense precipitation events along the Mediterranean coast. Natural Hazards and Earth System Sciences, 13, 1707- 1722.

Reboita, M.S. Ciclones Extratropicais sobre o Atlântico Sul: Simulação Climática e Experimentos de Sensibilidade. 359 f. Tese (Doutorado em Meteorologia) - Instituto de Astronomia, Geofísica e Ciências Atmosféricas – IAG – USP, São Paulo, 2008.

Reboita, M.S.; Crespo, N.M.; Torres, J.A.; Reale, M.; Rocha, R.P.; Giorgi, F.; Coppola, E., 2021.

Future Changes in winter explosive cyclones over the Southern Hemisphere domains from the CORDEX-CORE ensemble. Climate Dynamics, 57, 3303-3322.

Reboita, M.S.; Gozzo, L.F.; Crespo, N.M.;

Custodio, M.S.; Lucyrio, V.; De Jesus, E.M.;

Rocha, R.P., 2022. From a Shapiro-Keyser extratropical cyclone to the subtropical cyclone Raoni: An unusual winter synoptic situation over the South Atlantic Ocean. Quarterly Journal of the Royal Meteorological Society, 148, 1-19.

Reboita, M.S.; Rocha, R.P.; Ambrizzi, T.;

Suguhara, S., 2010. South Atlantic Ocean cyclogenesis climatology simulated by regional climate model (RegCM3). Climate Dynamics, 35, 1331-1347.

Reis, P.A.; Aquino, F.E.; Schossler, V.; Bernardo, R.T.; Simões, J.C., 2020. Tropical-Antarctic connections of an explosive cyclone in Southern Brazil: rainfall stable isotope ratios and atmospheric analysis. Advances in Polar Science, 31, 103-111.

Rienecker, M.M.; Suarez, M.J.; Gelaro, R.;

Todling, R.; Bacmeister, J.; Liu, E.;

Bosilovich, M.G.; Schubert, S.D.; Takacs, L.;

Gi-Kong, K.; Bloom, S.; Chen, J.; Collins, D.;

Conaty, A.; Silva, A.; Gu, W.; Joiner, J.;

Koster, R.D.; Lucchesi, R.; Molod, A.; Owens, T.; Pawson, S.; Pegion, P.; Redder; C.R.;

Reichle, R.; Robertson, F.R.; Ruddick; A.G.;

Sienkiewiks, M.; Woollen, J., 2011. MERRA:

(15)

2795

Andrade, H. N., Nunes, A. B., Alves, R.C.M.

NASA's Modern-Era Retrospective Analysis Research and Applications. Journal of Climate 24, 3624-3648.

Rudeva, I; Gulev, S., 2011. Composite analysis of North Atlantic extratropical cyclones in NCEP-NCAR reanalysis data. Monthly Weather Review, 139, 1419-1446.

Rogers, E.; Bosart, L.F., 1986. An investigation of explosively deepening oceanic cyclones, Monthly Weather Review, 114, 702-718.

Sanders, F., 1986. Explosive Cyclogenesis in the West-Central North Atlantic Ocean, 1981- 84.

Part I: Composite structure and mean behavior.

Monthly Weather Review, 114, 1781-1794.

Sanders, F., 1993. Notes and Correspondence.

Upper-level geostrophic diffluence and deepening of surface lows. Weather Forecast, 8, 339-344.

Sanders, F.; Gyakum J.R., 1980. Synoptic- dynamic climatology of the “Bomb”. Monthly Weather Review, 108, 1589-1606.

Santurette, P.; Georgiev, C.G. Weather analysis and forecasting: Applying satellite water vapor and potential vorticity analysis. Amsterdam:

Academic Press, 2005, 179 p.

Schemm, S.; Rudisuhli, S. Sprenger, M., 2020.

The life cycle of upper-level troughs and ridges: a novel detection method, climatologies and Lagrangian characteristics. Weather and

Climate Dynamics, 1, 459-479.

Seluchi, M.E.; Saulo, A.C., 2012. Baixa do Noroeste Argentino e Baixa do Chaco:

Características, diferenças e semelhanças.

Revista Brasileira de Meteorologia. Disponível em: https://www.scielo.br/pdf/rbmet/

v27n1/a06v27n1.pdf.

Seluchi, M.E; Saulo, A.C.; Nicolini, M.;

Satyamurty, P., 2003. The Northwestern Argentinean Low: a study of two typical events. Monthly Weather Review, 131, 2361- 2378.

Uccellini, L.W., 1986. The Possible Influence of Upstream Upper-Level Baroclinic Processes on the Development of the QE II Storm.

Monthly Weather Review 114, 1019-1027.

Uccellini, L.W.; Keyser D.; Brill K.F.; Wash C.H., 1985. The President’s Day Cyclone of 18-19 February 1979: Influence of upstream trough amplification and associated tropopause folding on rapid cyclogenesis. Monthly Weather Review, 113, 962-988

Wallace, J.M.; Hobbs, P.V. Atmospheric Science – An Introductory Survey. 2 ed. San Diego:

Elsevier, 2006. 483 p.

Wang, C.C.; Rogers, J.C., 2001. A composite study of explosive cyclogenesis in different sectors of the North Atlantic. Part I: Cyclone structure and evolution. Monthly Weather Review, 129, 1481-1499.

Wash, C.H.; Hale, R.A.; Dobos, P.H.; Wright, E.J., 1992. Study of explosive and nonexplosive cyclogenesis during FGGE.

Monthly Weather Review, 120, 40-51.

Whitacker, J. S.; Uccellini, L. W.; Brill, K. F.

1988. A Model-Based Diagnostic Study of the Rapid Development Phase of the Presidents’

Day Cyclone. Monthly Weather Review 116, 2337- 2365.

Willison, J.; Robinson, W.A.; Lackmann, G.M., 2013. The importance of resolving mesoscale latent heating in the North Atlantic Storm Track. Journal of the Atmospheric Sciences, 70, 2234-2249.

Zhang, S.; Fu, G.; Lu. C.; Liu, J., 2017.

Characteristics of Explosive Cyclones over the Northern Pacific. Journal of Applied Meteorology and Climatology, 56, 3187-3210.

Zhang Z.; Ralph, F.M.; Zheng, M., 2019. The relationship between extratropical cyclone strength and atmospheric river intensity and position. Geophysical Research Letters, 46, 1814-1823.

Referências

Documentos relacionados

Despercebido: não visto, não notado, não observado, ignorado.. Não me passou despercebido

Ao Dr Oliver Duenisch pelos contatos feitos e orientação de língua estrangeira Ao Dr Agenor Maccari pela ajuda na viabilização da área do experimento de campo Ao Dr Rudi Arno

Neste trabalho o objetivo central foi a ampliação e adequação do procedimento e programa computacional baseado no programa comercial MSC.PATRAN, para a geração automática de modelos

Ousasse apontar algumas hipóteses para a solução desse problema público a partir do exposto dos autores usados como base para fundamentação teórica, da análise dos dados

Extinction with social support is blocked by the protein synthesis inhibitors anisomycin and rapamycin and by the inhibitor of gene expression 5,6-dichloro-1- β-

didático e resolva as ​listas de exercícios (disponíveis no ​Classroom​) referentes às obras de Carlos Drummond de Andrade, João Guimarães Rosa, Machado de Assis,

i) A condutividade da matriz vítrea diminui com o aumento do tempo de tratamento térmico (Fig.. 241 pequena quantidade de cristais existentes na amostra já provoca um efeito

Peça de mão de alta rotação pneumática com sistema Push Button (botão para remoção de broca), podendo apresentar passagem dupla de ar e acoplamento para engate rápido