DISSERTATION zur Erlangung des Doktorgrades an der Naturwissenschaflichen Fakultät der Karl-Franzens-Universität Graz

Properties of the solar velocity field indicated by motions of groups and coronal bright points

Ivana Poljančić Beljan

begutachtet von

Univ.-Prof. Dr. Arnold Hanslmeier Institut für Physik Karl-Franzens-Universität Graz

und

Dr. Roman Brajša, scientific advisor Hvar Observatory, Faculty of Geodesy University of Zagreb

2018 Thesis advisor: Univ.-Prof. Dr. Hanslmeier Arnold Ivana Poljančić Beljan

Properties of the solar velocity field indicated by motions of sunspot groups and coronal bright points

Abstract

The solar dynamo is a consequence of the interaction of the solar magnetic field with the large scale plasma motions, namely differential rotation and meridional flows. The main aims of the dissertation are to present precise measurements of the solar differential rotation, to improve insights in the relationship between the solar rotation and activity, to clarify the cause of the existence of a whole range of different results obtained for meridional flows and to clarify that the observed transfer of the angular momentum to- wards the solar equator mainly depends on horizontal Reynolds stress. For each of the mentioned topics positions of sunspot groups and coronal bright points (CBPs) from five different sources have been used. Synodic angular rotation velocities have been calculated using the daily shift or linear least-square fit methods. After the conversion to sidereal values, differential rotation profiles have been calculated by the least-square fitting. Covariance of the calculated rotational and meridional velocities was used to de- rive the horizontal Reynolds stress. The analysis of the differential rotation in general has shown that Kanzelhöhe Observatory for Solar and Environmental research (KSO) provides a valuable data set with a satisfactory accuracy suitable for the investigation of differential rotation and similar studies. A significant correlation was found between the solar equatorial rotation speed and indices of the solar activity, for both tracers. Dif- ferent CBPs and sunspot groups results for the meridional flows are consistently inter- preted. The analysis of the horizontal Reynolds stress, by both tracers, showed transfer of angular momentum towards the solar equator, which is consistent with the observed solar differential rotation profile. The results presented in the thesis are original scien- tific contributions primarily fundamental in nature, but a part of the research also has possible applicative potential for the use of dynamo modelling.

ii Thesis advisor: Univ.-Prof. Dr. Hanslmeier Arnold Ivana Poljančić Beljan

Eigenschaften des solaren Geschwindigkeitsfeldes angegeben durch Sonnenfleckengruppen und koronale helle Punkte

Abstract (German)

Der Sonnendynamo ist eine Folge der Wechselwirkung des solaren Magnetfeldes mit großräumigen Plasmabewegungen, die sich durch differentielle Rotation und merid- ionale Bewegung zeigt. Daher ist das Ziel dieser Dissertation, präzise Messungen der differentiellen Rotation der Sonne zu präsentieren, die Abhängigkeit von Rotation und Aktivität besser zu verstehen, die Ursache für die Existenz verschiedener Resultate für meridionale Strömungen zu erklären und die Abhängigkeit der erwähnten Bewegung durch das Drehmoment und der Strömung in Richtung des Sonnenäquators mit der hor- izontalen Reynoldsspannung zu klären. Für jedes der genannten Themen wurden Posi- tionen von Sonnenfleckengruppen und hellen koronalen Punkten (CBPs) von fünf ver- schiedenen Quellen verwendet. Synodische Umlaufzeiten wurden nach der zweitägigen Abstand und der Methode der kleinsten Quadrate berechnet und ausgewertet. Nach Um- rechnung der synodischen Umlaufzeiten in die siderischen Umlaufzeiten, wurden dif- ferentielle Rotationsprofile unter Berücksichtigung der Methode der kleinsten Quadrate erstellt. Die Kovarianz der ermittelten Rotations- und Meridionalgeschwindigkeiten wurde verwendet, um die horizontale Reynoldsspannung zu berechnen. Die Unter- suchung der differentiellen Sonnenrotation hat gezeigt, dass der Datenbestand des KSO mit seiner zufriedenstellenden Genauigkeit für die Analyse der differentiellen Rotation und ähnlicher Studien durchaus geeignet ist. Es wurde eine statistisch bemerkenswerte Korrelation zwischen der Rotationsgeschwindigkeit am Äquator und dem Indikator der Sonnenaktivität für die beiden Tracer festgestellt. Die Analyse der horizontalen Reynoldsspannung beider Tracer zeigte auch eine Ablenkung des Drehimpulses in Richtung des Sonnenäquators. Die in dieser Dissertation dargestellten Ergebnisse sind ursprüngliche wissenschaftliche Beiträge mit fundamentalem Charakter, während ein Teil der Forschung das Potential besitzt, das Model des Dynamoprozesses besser zu beschreiben.

iii Acknowledgements

Special thanks go to my supervisor Univ.-Prof. Dr. Arnold Hanslmeier and my mentor Dr. Roman Brajša. You were always there, opened for conversation and giving answers, encouraging me to go constantly forward. I am deeply grateful to Univ.-Prof. Dr. Rajka Jurdana-Šepić for being there in every moment, for responding to my e-mails, always in an amazingly short time, for all the sleepless working nights with thousands of e-mails. I would like to thank Dr. Werner Pötzi from the Kanzelhöhe Observatory for So- lar and Environmental Research for the hospitality during my stay at Kanzelhöhe, for patience and understanding during our long conversations, for giving answers to many questions concerning the topics connected with the Kanzelhöhe sunspot drawings and full disc white light CCD images. I also thank to Assoc. Univ.-Prof. Dr. Astrid Veronig for discussions during my attendance on the PhD seminars, which prompted me to deepen and broaden the analysis of the Kanzelhöhe data. I am deeply grateful to Mag. Is- abell Piantschitsch for being my guide in solving the whole variety of paperwork during my doctoral studies and for being so kind and friendly. Without most of the people from Hvar Observatory this thesis would not have be- come what it is. I thank Dr. Bojan Vršnak. Special thanks goes to Dr. Davor Sudar and Dr. Domagoj Ruždjak for the fruitful discussions and plenty of suggestions. Dr. Ivica Skokić helped me immensely, by giving plenty of advice concerning the programming in Python. I thank M.Sc.M.E. Damir Hržina from Zagreb Observatory for his assistance during the work on KSO paper in upgrading the program ”Sungrabber” according to my requirements, whenever I needed. This thesis was financially supported in part by Croatian Science Foundation under the project 6212 “Solar and Stellar Variability” and in part by the University of Rijeka under the project number 13.12.1.3.03. Without these financial supports my travels to Graz would be quite difficult. There are no words that can describe my gratitude to my husband, his parents and my parents, especially to my mother. Thank you mom for suporting me in all aspects, espe- cially by looking after my son whenever I needed. I still feel your support, somewhere from above.

iv Contents

1 Introduction 1 1.1 List of publications ...... 6

2 Observations of the solar differential rotation 8 2.1 and sunspot groups as tracers ...... 10 2.1.1 Magnetic properties of sunspots ...... 10 2.1.2 The evolution of a sunspot ...... 12 2.2 Coronal bright points as tracers ...... 14 2.3 Surface rotation ...... 15 2.4 Internal rotation ...... 17 2.5 Meridional motions ...... 19 2.6 Horizontal Reynolds stress ...... 21 2.7 Relationship between the solar rotation and activity ...... 23

3 Theory of the solar differential rotation 26 3.1 Angular momentum transport and magnetic braking ...... 26 3.2 Axisymmetric and non-axisymmetric convective motions ...... 27 3.3 The Λ-effect and eddy viscosity ...... 28 3.4 Angular momentum transport ...... 31 3.5 Models of the differential rotation ...... 32 3.5.1 Mean-field models ...... 33 3.5.2 Global (explicit) models ...... 34

4 Peer-reviewed publications 36 4.1 Paper I: Solar differential rotation in the period 1964 - 2016 determined by the Kanzelhöhe data set ...... 36 4.2 Paper II: A relationship between the solar rotation and activity in the pe- riod 1998–2006 analysed by tracing small bright coronal structures in SOHO-EIT images ...... 48 4.3 Paper III: A Relationship between the Solar Rotation and Activity Anal- ysed by Tracing Sunspot Groups ...... 55 4.4 Paper IV: Meridional motions and Reynolds stress from SDO/AIA coro- nal bright points data ...... 67

v 4.5 Paper V: Meridional Motion and Reynolds Stress from Debrecen Photo- heliographic Data ...... 74

5 Conclusion 88

References 96

vi 1 Introduction

Solar activity manifests itself in many diverse phenomena (sunspots, flares, coronal mass ejections, etc.) that vary on several distinct timescales. The most prominent manifestation of solar activity is the 11-year Schwabe cycle. Extensive reviews on the long-term behaviour of the solar activity and the Schwabe cycle are given by Hathaway (2010) and Usoskin (2017). It is generally accepted that the is magnetic in nature and generated by dynamo processes within the . The Sun’s magnetic field is maintained by its interaction with plasma motions, i.e. by the differential rotation of the convective zone. The physical processes underlying the complex magnetohydrodynamical system are still not fully understood, and a large number of dynamo models have been considered (Cameron et al. 2017). Precise measurements of the solar differential rotation, its variations, and its correlation with the solar activity give important observational constraints on the modelling of the solar dynamo. Solar rotation can be generally determined with one of the three methods: the tracer method, the spectroscopic method and the helioseismology method. With the tracer method, used in this work, the solar rotation is measured by following postitions of the structures located in the solar atmosphere in time. Due to the feature’s physical background and the conditions in the layer of the solar atmosphere, where these features are located, different results can be found, which means that each layer has its own rotation law. Coronal bright points (CBPs) and sunspot groups turned out to be good tracers for the determination of coronal and photospheric differential rotation, respectively. These tracers complement each other and are therefore both used as tracers in this PhD thesis. Furthermore, coronal differential rotation determined by CBPs shows a good agreement with the photospheric differential rotation determined by sunspot groups.

1 Table 1.1: Research topics and tracers used in the peer-reviewed publications included in the thesis (Paper I - Paper V from the List of publications)

Research topic Sunspot groups as tracers CBPs as tracers Differential rotation in Paper I (KSO) - general Paper III (GPR, Rotation vs. activity Paper II (SOHO/EIT) USAF/NOAA, DPD) Transport of angular Paper V (DPD) Paper IV (SDO/AIA) momentum

The main aim of this PhD thesis is to investigate the properties of the solar velocity field (rotational and meridional flows, their correlation and covariance, Reynolds stress) indicated by motions of the CBPs and sunspot groups in order to reveal the transfer of angular momentum from higher to lower latitudes to maintain the observed differential rotation profile. The relationship between solar rotation and activity is also investigated. Complete publication list of the author is available at the end of this section. The thesis includes research contained in five peer-reviewed publications, covering topics previously mentioned and published in Astronomy and Astrophysics and Solar Physics journals that belong to the Q1 and/or Q2 quartiles according to Scimago Journal Rankings (Paper I - Paper V from the List of publications) - Table 1.1. A few more papers (Paper IX, X and XI from the List of publications) had preceded the analysis contained in this thesis. First of them, Paper IX - Poljančić et al. (2010), checked the precision of the Solar Optical Observing Network/United States Air Force/National Oceanic and Atmospheric Administration (SOON/USAF/NOAA) data set by comparing it with the Greenwich Photoheliographic Results (GPR) data set. The Kanzelhöhe Observatory for Solar and Environmental Research (KSO) data set (heliographic coordinates determined from the sunspot drawings for the years 1972 and 1993) has been used as a reference for the comparison of sunspot position measurements. It was found that the SOON/USAF/NOAA data are somewhat less accurate than the GPR ones. The second paper, Paper X - Poljančić et al. (2011), expanded that analysis on several more data sets, again by using the KSO data set as a reference, for the comparison of sunspot position measurements. It turned out that, in almost all cases, the latitude, longitude and synodic angular velocity differences between the KSO and other data sets are the smallest for Debrecen Photoheliographic Data (DPD), GPR and

2 SOON/USAF/NOAA. In the third paper, Paper XI - Poljančić Beljan et al. (2014), differential rotation parameters are calculated and differential rotation profiles for the solar cycles nos. 20 and 22 are shown, for the KSO data in preliminary form. The GPR and SOON/USAF/NOAA data sets have already been widely used for the analysis of differential rotation in general as well as for the analysis of the transport of the angular momentum (meridional motions, horizontal Reynolds stress) and relationship between the rotation and activity. Following the conclusion in Paper X- Poljančić et al. (2011) mentioned above, that latitude, longitude and synodic angular velocity differences between the KSO and DPD, KSO and GPR, and KSO and SOON/USAF/NOAA are the smallest, we have noticed that it would be of interest to process the DPD and KSO data sets. Therefore, three articles have emerged from this idea, related to the processing of the KSO and DPD sunspot groups data and dealing with three research topics - differential rotation in general, relationship between the rotation and activity and transport of the angular momentum discussed through meridional motions and horizontal Reynolds stress (Paper I, Paper III, Paper V). Paper II and Paper IV represent a kind of complement to that research, performed using another tracers - CBPs. All of the above is summarized in Table 1.1. Details, concerning each paper separately, are available hereafter.

Paper I (Sec. 4.1) - Poljančić Beljan et al. (2017)

The main aim of this work is to determine the solar differential rotation by tracing sunspot groups during the period 1964 - 2016, using the KSO sunspot drawings and white light images. Until now, solar differential rotation has been investigated only for the period 1947-1981 with the KSO data, so the main aim was to provide a continuation of this analysis. For this reason KSO sunspot drawings for solar cycles nos. 20 - 23 (interactive method – using a special software Sungrabber) and KSO white light images for solar cycle no. 24 (automatic method – the algorithm for automatic recognition of the sunspot groups), covering half a century, were processed by the author. In this paper we calculate differential rotation parameters for the whole analysed period, as well as for the each cycle separately, and compare them with the differential rotation parameters collected from other data sets. An investigation of the north - south rotational asymmetry is also performed and a barely noticeable north - south asymmetry is observed for the whole 1964 - 2016 time period. It is shown that the KSO data set used in this paper is in a good agreement with the DPD and GPR data sets and is suitable for the investigation of the long term variations of the solar rotation profile. As a continuation of this work an analysis of meridional motions and horizontal Reynolds stress from KSO is already performed (Ruždjak et al. 2018) - Paper VIII from

3 the List of publications. Investigation of the temporal variation of the solar rotation parameters (cycle to cycle variations, variations in single years, variations during the solar cycle) and the relationship between the parameters of the solar rotation and the solar activity using the yearly values of the solar rotation parameters derived from the KSO data, which has never been used before for this purpose, is going to be the third paper in this series.

Paper II (Sec. 4.2) - Jurdana-Šepić et al. (2011)

This paper is dealing with the full-disc solar images obtained with the Extreme ultraviolet Imaging Telescope (EIT) on board the Solar and Heliospheric Observatory (SOHO). The full-disc solar images obtained with that instrument were used to analyse solar differential rotation determined by tracing CBPs. Precise measurements of the solar differential rotation enabled a comparison of the determined solar rotation parameters with the status of solar activity in the 1998-2006 period. So, this study aims to find a relationship between the rotation of the CBPs described by the solar rotation parameters and indices of solar activity on monthly and yearly temporal scales.

Paper III (Sec. 4.3) - Ruždjak et al. (2017)

In this paper the sunspot groups positions from the GPR, the USAF/NOAA and the DPD data sets in the period from 1874 to 2016 were used to calculate yearly values of the solar differential rotation parameters A and B. These differential rotation parameters were compared with the indices of the solar activity (yearly mean total sunspot numbers), so a relationship between solar activity and differential rotation was examined. Both secular and cycle-related changes of the equatorial rotation rate and differential rotation parameter B were investigated. Also, the equatorial rotation velocity during the minimum of activity was reviewed and correlation between the peak height of the equatorial rotation velocity and the strength of the next maximum in comparison to the previous one, was observed.

Paper IV (Sec. 4.4) - Sudar et al. (2016)

In this paper, Solar Dynamics Observatory/Atmospheric Imaging Assembly (SDO/AIA) images are used to measure solar differential rotation profile very accurately and to investigate meridional flows, torsional oscillations, and horizontal Reynolds stress based on approximately 6 months of CBP data. It is shown that meridional motion is predominantly poleward for all latitudes, while solar velocity residuals show signs of torsional oscillations. Horizontal Reynolds stress, as an indicator for explaining

4 the observed solar rotation profile, was found to be smaller than in similar works, but still showed transfer of angular momentum towards the solar equator, as expected from observations.

Paper V (Sec. 4.5) - Sudar et al. (2017)

The main focus of this paper is to analyse the meridional motion, rotation velocity residuals and horizontal Reynolds stresses (transfer of angular momentum towards the solar equator), by tracing sunspot groups over the solar surface using the DPD data in the 1974 – 2016 period. Apart from the standard differential rotation profile, meridional motion directed towards the zone of solar activity was found. Also, a statistically significant dependence of meridional motion on rotation velocity residuals was found, confirming the transfer of angular momentum towards the equator. The analysis of horizontal Reynolds stress reveals that the transfer of angular momentum is stronger with increasing latitude up to about 40◦, where there is a possible maximum in absolute value.

The thesis is compound of 5 sections, Sec. 1 being an overview of the thesis. In the Sec. 2 an insight to the observations of the solar differential rotation, closely connected with the investigations in the thesis, is given: physical background of the used tracers - sunspot groups and coronal bright points, overview of the previous observations of the surface and internal rotation, meridional motions and transport of the angular momentum, i.e. horizontal Reynolds stress, as well as the relationship between the rotation and activity. In Sec. 3 an overview of the theory of the solar differential rotation is given, based on the derivation of the angular momentum conservation law, which is crutial for understanding the observed transport of the angular momentum towards the solar equator. The results are presented in Sec. 4 through 5 included peer - reviewed publications. Conclusions are drawn in Sec. 5.

5 1.1 List of publications

Peer-reviewed publications

I Poljančić Beljan, I., Jurdana-Šepić, R., Brajša, R., Sudar, D., Ruždjak, D., Hržina, D., Pötzi, W., Hanslmeier, A., Veronig, A., Skokić, I., Wöhl, H.: So- lar differential rotation in the period 1964 - 2016 determined by the Kanzelhöhe data set, Astronomy and Astrophysics, 606 (2017), A72

II Jurdana-Šepić, R., Brajša, R., Wöhl, H., Hanslmeier, A., Poljančić, I., Svalgaard, L., Gissot, S. F.: A relationship between the solar rotation and activity in the period 1998–2006 analysed by tracing small bright coronal structures in SOHO- EIT images, Astronomy and Astrophysics, 534 (2011), A17

III Ruždjak, D., Brajša, R., Sudar, D., Skokić, I., Poljančić Beljan, I.: A Rela- tionship between the Solar Rotation and Activity Analysed by Tracing Sunspot Groups, Solar Physics, 292 (2017), 179

IV Sudar, D., Saar, S. H., Skokić, I., Poljančić Beljan, I., Brajša, R.: Meridional mo- tions and Reynolds stress from SDO/AIA coronal bright points data, Astronomy and astrophysics, 587 (2016), A29

V Sudar, D., Brajša, R., Skokić, I., Poljančić Beljan, I., Wöhl, H.: Meridional Mo- tion and Reynolds Stress from Debrecen Photoheliographic Data, Solar Physics, 292 (2017), 86

VI Poljančić Beljan, I., Jurdana-Šepić, R., Semkov, E. H., Ibryamov, S., Peneva, S. P., Tsvetkov M. K.: Long-term photometric observations of pre-main se- quence objects in the field of North America/Pelican Nebula, Astronomy and As- trophysics, 568 (2014), A49

VII Vršnak, B., Dumbović, M., Čalogović, J., Verbanac, G., Poljančić Beljan, I., Geomagnetic effects of corotating interaction regions, Solar Physics, 292 (2017), 140

VIII Ruždjak, D., Sudar, D., Brajša, R., Skokić, I., Poljančić Beljan, I., Jurdana-Šepić, R., Hanslmeier, A., Veronig, A., Pötzi, W., Meridional Motions and Reynolds Stress Determined by using Kanzelhöhe Drawings and White Light Solar Images from 1964 to 2016, Solar Physics, 293 (2018), 59

6 Other publications

IX Poljančić, I., Brajša, R., Ruždjak, D., Hržina, D., Jurdana-Šepić, R., Wöhl, H., Otruba, W.: A comparison of sunspot position measurements from different data sets, Sun and Geosphere, 5 (2010), 52-57

X Poljančić, I., Brajša, R., Hržina, D., Wöhl, H., Hanslmeier, A., Pötzi, W., Baranyi, T., Özgüç, A., Singh, J., Ruždjak, V.: Differences in heliographic posi- tions and rotation velocities of sunspot groups from various observatories, Cent. Eur. Astrophys. Bull., 35 (2011), 59–70

XI Poljančić Beljan, I., Jurdana-Šepić, R., Čargonja, M., Brajša, R., Hržina, D., Pötzi, W., Hanslmeier, A.: An analysis of the solar differential rotation from the Kanzelhöhe sunspot drawings, Cent. Eur. Astrophys. Bull., 38 (2014), 87–93

XII Jurdana-Šepić, R., Poljančić Beljan, I.: A search for T Tauri stars and related objects: archival photometry of candidate variables in V733 Cep field, Cent. Eur. Astrophys. Bull., 39 (2015) 1, 87-93

XIII Skokić, I., Sudar D., Saar, S. H., Brajša, R., Poljančić Beljan, I.: An application of the random walk model to proper motions of coronal bright points from SDO data, Cent. Eur.Astrophys.Bull., 40 (2016) 1, 23–28

7 2 Observations of the solar differential rotation

The hydrogen-burning core occupies roughly the inner third of the Sun’s radius, the radiative zone the next third and the convection zone the outher third of the radius. The convection zone terminates just below the photosphere, a few hundred kilometres thick layer of the atmosphere which can be seen in the visible portion of the continuous radiation spectrum (Wilson 1994; Thomas & Weiss 2008). Long term observational studies of the large-scale motions in the photosphere and convection zone play a major rule in shaping our understanding of the activity cycle. Large-scale motions imply velocity fields observed at the solar surface on a scale larger than supergranulation, like differential rotation and meridional flows, extensively re- viewed in this thesis. Large-scale motions have been reviewed by Howard (1978, 1984), Schröter (1985), Bogart (1987), who concentrated on the solar rotation, and Beckers (1981), who discussed the overall dynamics of the photosphere. The studies of these motions and their patterns of change in the Sun’s interior became possible with the help of helioseismology. More precisely, there are three categories of observations of large- scale motion patterns: the tracer method, the spectroscopic method (measurements of Doppler shifts) and the helioseismology method (Beck 2000). The first two categories are used for determination of patterns of motion at the solar surface and measurements are covering a period of many years (half a century or more), while helioseismology deals with the Sun’s interior and covers a shorter time period. Details of the propaga- tion of the seismic waves through the Sun, inferred from the resonant frequencies, reveal the Sun’s inner structure and the internal angular velocity (Duvall et al. 1984; Schou et al. 1998). On the other hand, refining the instruments and techniques for the reduc-

8 tion of the Doppler and tracer data is also of interest and has been rapidly progressed. The study of patterns of motion associated with the solar cycle requires the analysis of data collected over a time comparable to the cycle period or longer, which can not be covered by helioseismology (Beck 2000). Interpretation of lineshift data requires complex analysis (removing of the motion components due to the Earth’s rotation and orbit, removing of the instrumental drift, stray light), while tracer measurements are relatively straigthforward, with a smaller number of systematic errors present. On the other hand, the random noise level for tracer measurements is one or two orders of magnitude greater than it is for Doppler data. The source of the random noise problem in the tracer measurements is determination of the average velocities during the time between tracer observations, while Doppler shifts measure instantaneous velocity. Another fundamental difference between the Doppler and tracer measurements is line of sight components of the solar velocities given by Doppler shifts but transverse components given by tracers (Schröter 1985). Most of the objects, which are usually used as tracers, consist of bundles of flux tubes. These tubes stretch radially outwards and show the rotation of deeper magnetic layers where their footpoints are located (see Ruždjak et al. (2004)). On the other hand, Doppler shifts show the rotation of the currently observed layer. This is discussed in more details later (see Fig. 2.6). With the tracer method, the solar velocity fields are measured by following positions of the structures located in the solar atmosphere in time, like sunspots, or by correlation methods, like the magnetic features in a magnetogram. Typical tracers for the photo- sphere are sunspots and sunspot groups, faculae, magnetic field patterns, for the chro- mosphere Ca- and Hα -filaments and for the corona EUV plages, green- and white-light structures, e.g. CBPs (Howard 1984). Moreover, not all structures are suitable to be used as tracers for measurements of the solar rotation. Coronal bright points (CBPs) and sunspot groups turned out to be good tracers for the determination of coronal and photospheric differential rotation, re- spectively. Sunspots and sunspot groups are very often used as tracers, mainly because of the long term data sets available from various observatories. CBPs are usually stud- ied during the short time periods, for which the satellite data is available, but are well distributed all over the solar surface, even in polar regions where sunspots do not appear. Hence, these tracers complement each other. Furthermore, coronal differential rotation determined by CBPs shows a good agreement with the photospheric differential rota- tion determined by sunspot groups, when the height corrections of CBPs are taken into account (Brajša et al. 2004; Wöhl et al. 2010). These conclusions became an inspiration for starting analysis of the solar differential rotation using sunspot groups and CBPs as tracers, presented in the thesis.

9 2.1 Sunspots and sunspot groups as tracers

Sunspots and sunspot groups have been used as tracers for solar rotation from the time they were first recognized as features on the Sun, up to the present day. Numerous studies of solar rotation, performed with sunspot groups as tracers, exist in literature dating from Scheiner (1630), through Carrington (1863), Newton & Nunn (1951) to the huge list of recent studies presented in Table 3 of Paper I (Sec. 4.1). In three peer- reviewed publications included in the thesis, Paper I, III and V, sunspot groups have been used as tracers.

2.1.1 Magnetic properties of sunspots Magnetic activity on the Sun is not uniformly distributed over the solar surface, but instead is concentrated into active regions where sunspots represent an essential ingre- dient. Binding of pores and smaller magnetic flux tubes results in a single, growing pore which represents the umbra (Vrabec 1974; Zwaan 1978, 1992; McIntosh 1981). If a diameter of a pore reaches a sufficient size of about 3500 km or more, or sufficient total magnetic flux of order 1 × 1020 Mx, a penumbra is formed (Leka & Skumanich 1998).

Figure 2.1: Theoretical models of the magnetic field beneath the visible surface of the umbra: a) „monolithic model“, b) „cluster model“, c) detailed „cluster model“. Parts a) and b) are from Thomas & Weiss (2008), c) is from Parker (1979).

The darkness of sunspots may be interpreted by strong magnetic fields that partially inhibit the usual transport of energy by convection at the solar photosphere or just below it. A radius of 10 000 - 20 000 km is characteristic for a well developed spot, where a dark central nucleus represents the umbra which is surrounded by a less dark, fila- mentary penumbra. At the centre of the spot, the magnetic field strength is typically around 2800 G and it is almost vertical. Increasing the radius, the inclination of the

10 Figure 2.2: Sunspot image with the visible umbral dots, indicating convection in the umbra. The image was made by the Swedish 1-m Solar Telescope (SST) that is operated on the island of La Palma by the Institute for Solar Physics from Stockholm, Sweden.* magnetic field to the vertical is also increasing, where the field strength is lower, reach- ing some 1000 G or less (Wilson 1994). The energy flux through a given area of the umbra is ∼ 20% of that through an equivalent area of the photosphere, corresponding to a temperature deficit of 2000 K, while the average penumbral intensity is about 75% of that outside the spot, corresponding to a deficit of only 400 K (Thomas & Weiss 2008; Wilson 1994; Stix 2002). Biermann (1941) offered the standard explanation of this cooling – the convective transport of energy below the spot is inhibited by its underlying strong magnetic field. But, the calculations that followed have shown that the surface energy flux would be reduced to only ∼ 5% of the photospheric flux, if inhibition of convection by the field below a spot is assumed. So, Parker (1979) introduced the „cluster model” in which the sub-umbral magnetic field structure is divided into bundles of flux, separated by non- magnetic regions (Fig. 2.1, b) and c)). In the „cluster model”, formation of umbral dots is explained by energy transport that is unimpeded in the field-free regions. Umbral dots are clearly visible on Fig. 2.2. Another model of the magnetic field of the umbra, a ”monolithic model”, consists of a monolithic flux tube (part a) of the Fig. 2.1). In order to observe a dark network enclosing bright umbral dots, a monolithic magnetic field should rule out oscillatory convection to reach the surface (Thomas & Weiss 2008). Unlike the umbra, the sunspot penumbra shows a strongly magnetized medium

*Observations were performed on 22nd Aug 2003 by Oddbjorn Engvold, Jun Elin Wiik and Luc Rouppe van der Voort. The image was downloaded from http://www.isf.astro.su.se/.

11 Figure 2.3: The interlocking-comb structure of the magnetic field in the filamentary penumbra of a sunspot. From Thomas & Weiss (2008) around the individual flux tube. Both the strength and the direction of the magnetic field has to be taken into account at the discontinuities at the surface of the flux tube, which requires complex surface currents. Since 1990, a large number of papers have confirmed, using the high-resolution observations, the interlocking-comb structure of the magnetic field. Fig. 2.3 shows the interlocking-comb structure of the magnetic field in the filamentary penumbra. Two kinds of filaments appear in penumbra: the bright fil- aments, where the magnetic field at the umbra-penumbra boundary is inclined at 30◦ and in the outher penumbra at 50◦–60◦ to the vertical, and the dark filaments, where the mag- netic field at the umbra-penumbra boundary is inclined at 65◦ and in the outer penumbra is almost horizontal. Magnetic flux tubes of the dark filaments typically spread into a shallow canopy lying just above the solar surface, while some of them dive back below the surface. On the other hand, flux tubes emerging from bright filaments emerge into the corona and extend over great distances across the Sun (Thomas & Weiss 2008).

2.1.2 The evolution of a sunspot The „cool cone” model of the development of a sunspot was presented by Wilson (1968). The first phase of the development, just before the appearance of the pore, is shown in Fig. 2.4 (left part). The deep convection is already inhibited by the strong twisted field and the cool region is developed. At the same time, the horizontal transfer of energy by radiation and convection decreases the diameter of the cool region in the upper part of the convection zone and a „cool cone” is generated. As long as the magnetic field is sufficiently deep, the surface granules appear bright as the „cool cone” does

12 Figure 2.4: The evolution of a sunspot. From Wilson (1968) not reach them. The flux tubes are shown with dashed lines, and in this stage are still horizontally directed, but are beginning to arch upwards. In the next step (Fig. 2.4, right part) the field becomes more twisted, causing a rise of buoyancy and forcing the flux tubes to become more arch-shaped. A „cool cone” finally reaches the surface, disturbing the visible granules by cutting them off from their main source of thermal energy and causing the appearance of a pore. Intersection between the „cool cone” and the surface increases, causing the appearance of a penumbra, by twisting the granules with strengthened „frozen-in” field that surround the pore and transforming them into filaments (Wilson 1968). Simple equilibrium models, which suggest that the configuration of a pore becomes unstable when the inclination of the magnetic field at the edge of a pore increases to a critical value, causing the development of a penumbra, are given in Simon & Weiss (1970); Rucklidge et al. (1995); Hurlburt & Rucklidge (2000). According to the obser- vations of Martinez Pillet (1997), the critical inclination angle to the vertical is about 35◦. There are a few more papers dealing with a nonlinear development of the instability between the flux tube of a growing pore and surrounding convectively driven filamen- tary perturbations (Hurlburt et al. 2000; Thomas et al. 2002; Tildesley 2003; Tildesley & Weiss 2004; Weiss et al. 2004). This instability can produce the rudimentary penum-

13 bra, but there is then a sudden jump associated with the appearance of a fully developed penumbra with its interlocking-comb magnetic field configuration (Thomas & Weiss 2008).

2.2 Coronal bright points as tracers

Sunspot groups turned out to be good tracers for the determination of photospheric differential rotation because of the long term data sets available from various observatories, but they appear only on latitudes 35-40 deg. Modern measurements enabled a determination of the overall differential rotation, for all latitudes and layers of the Sun, using different structures. CBPs are well distributed all over the solar surface, even in polar regions where sunspots do not appear. Hence, CBPs complement sunspot groups and were, therefore, also used as tracers for analysis presented in the thesis (Paper II and Paper IV). CBPs are often called X-ray bright points (XBPs) as they appear like points in X-rays. There are two types of CBPs according to the positions of development: a) ephemeral active region CBPs that are developed within the same latitudes as the regular active regions and b) quiet Sun CBPs developed uniformly over the solar surface (Sattarov et al. 2002; McIntosh & Gurman 2005). CBPs are identified as a small scale magnetic features (a size of nearly 6·103 −20·104 km), with a lifetimes of about several minutes to a few days, the average lifetime being 8-9 hours (Priest et al. 1994; Brajša et al. 2008). Increased coronal temperature results in a decline of the total number of CBPs on the solar disk, which supports their position in the low corona (Sattarov et al. 2002; McIntosh & Gurman 2005), confirmed by Brajša et al. (2004), as well as in Karachik et al. (2006). Brajša et al. (2004) estimated the height of CBPs at 8000–12000 km above the photosphere. The Converging Flux Model clarifies the cause of the CBPs as reconnection of the emerging flux with an overlying coronal magnetic field (Priest et al. 1994). Fig. 2.5 shows three phases of the evolution of CBPs: the pre-interaction phase, the interaction phase and the cancellation phase. In the pre-interaction phase two magnetic fragments with opposite polarities begin to approach one another (i), until they are not sufficiently close to form the line of magnetic neutrallity, so called null-point (ii). In the interaction phase, in stage (iii) the null-point becomes CBP and moves upwards, while during stage (iv) hot plasma loops are produced and CBP starts to move downwards. In the cancellation phase, in stages (v) and (vi) the fragments meet in the photosphere as a cancelling magnetic feature (CMF) and CBP disappears.

14 Figure 2.5: The Converging Flux Model for an CBP developed by Priest et al. (1994). Stages in the approach and interaction of two magnetic fragments with opposite polarities are shown. In stages (i) and (ii), the field line that separates magnetic fragments is shown with dashed curves.

2.3 Surface rotation

The solar rotation is differential and lowers with increasing latitude, whereas both hemispheres can be treated as roughly symmetric in rotation. Carrington (1863) was 7 the first to formulate this conclusion mathematically in the form Ω ∼ sin 4 ψ, where ψ is the heliographic latitude. As pointed out by Thomas & Weiss (2008), this can be considered as a discovery of the differential rotation, that provided direct evidence of the fluid nature of the Sun, at least in its outer layers. Observational results concerning the sidereal angular velocity at the solar surface are very often described by the expression

Ω = A + B sin2 b + C sin4 b (2.1)

15 Figure 2.6: The differential rotation profiles calculated by different authors with Doppler and tracer meth- ods: photospheric Doppler (light solid curve), coronal features (long dashes), magnetic structures (dot- dashed), chromospheric features (short dashes), photospheric faculae (dots) and sunspots (heavy solid). From Howard (1984). where Ω is the angular rotation velocity, b is the heliographic latitude and the parameters A, B and C are obtained by the least-square fit to the measurements (Howard et al. 1980). The parameter A represents the equatorial rotation velocity, while B (as a negative value) represents gradient of the differential rotation, e.g. the difference in angular velocity between the pole and the equator. In other words, B specifies the degree of rotational nonuniformity. The parameter A very often shows standard deviation of the order of 0.01 deg/day, while parameters B and C show the order of 0.1 deg/day. When sunspots are used as tracers, parameter C can be neglected, as sunspots do not occur at latitudes higher than ≈ 40◦ (Stix 1989, 2002). As the functions sin2n b, n = 1, 2, ..., are not orthogonal, it is possible to get a crosstalk between the parameters B and C when eq. 2.1 is used. This problem has been discussed by many authors, e.g. by Howard et al. (1980), and then solved by Snodgrass (1984) who concluded that it is better to use a set of functions that are orthogonal on the solar disc. Therefore, Snodgrass (1984) introduced a representation by the Gegenbauer polynomials. Shortly after, Snodgrass & Howard (1985) clarified that use of Gegen-

16 bauer polynomials instead of eq. 2.1 is particularly important when the results of the solar rotation profile from different sources are compared. Regardless of this fact, from 1980’s onwards, eq. 2.1 has been used in a large number of papers. For instance, a long list of differential rotation profiles collected from different sources, using only sunspots and sunspot groups as tracers, is presented in Table 3 and shown in Fig. 5 of Paper I (Sec. 4.1). If the differential rotation profiles concerning Doppler and tracer data are compared (Fig. 2.6), they show a slight difference between each other, as well as the different tracers. Namely, due to a tracer’s physical background and the conditions in the layer of the solar atmosphere, where these tracers are located, different results concerning the solar differential rotation can be found, which means that each layer has its own rotation law. But, still, from Fig. 2.6 it is obvious that coronal features (long dashes) and sunspots (heavy solid) show very similar rotation profiles, with equatorial rotation velocities practically the same. This conclusion has been approved by comparing coro- nal bright points and sunspot groups rotation profiles in Brajša et al. (2004), Wöhl et al. (2010) and in Paper II (Sec. 4.2), included in the thesis.

2.4 Internal rotation

The decrease of angular velocity from the equator to poles seen on the solar surface is also present inside the convection zone, down to the tachocline, as helioseimology showed (Wilson et al. 1997; Schou et al. 1998). As can be seen in Fig. 2.7, the rota- tional nonuniformity extends throughout the convection zone down to the tachocline, pointing to the uniform rotation in the radiative zone (Kosovichev et al. 1997). Thus, the differential rotation and convective regions almost coincide, which goes in favor of theoretical assumptions about the origin of the differential rotation, explained by the interaction between convection and rotation. The angular velocity within the convection zone has approximately the same latitude dependence as at the surface, except for the layer between the solar surface and 0.95 r⊙, where some small but important changes are visible (Fig. 2.7). The solar tachocline represents the transition region between the convective shell and radiative core, and is marked with the vertical line on the Fig. 2.7. The radiative core (to the left of the vertical line) rotates almost uniformly, with the angular velocity which almost coincides with the one observed at the surface, at heliographic latitudes 30◦. Bertello et al. (2000) showed that the rotation frequency stays almost the same in the inner and outer parts of the core, but shows a significant increase of the error bars for r < 0.1r⊙ (Stix 2002). Fig. 2.8. presents the same as Fig. 2.7, depth profiles of the angular velocity for sev- eral latitudes, but developed with the mean-field model from Kitchatinov & Olemskoy

17 Figure 2.7: Angular velocity expressed by νrot = Ω/2π as a function of the relative radius for three helio- ◦ ◦ ◦ graphic latitudes, 0 , 30 and 60 derived with the Michelson Doppler Imager on SOHO. The shaded regions indicate the errors, the vertical line indicate the position of the base of the convection zone, while arrows mark the angular velocity rates measured on the solar surface. From Kosovichev et al. (1997)

Figure 2.8: The same as in Fig. 2.7, but here derived theoretically by the mean-field model from Kitchatinov & Olemskoy (2011)

(2011). A good agreement between theoretical and observational results is obvious. The change of the angular velocity at the solar tachocline (near r = 0.7r⊙) confirms the basic theoretical assumption that differential rotation is maintained by convection. Also, the solar magnetic field and 22-year magnetic cycle are consequences of this im- portant fact, and are closely connected with the solar dynamo throught the good agree-

18 ment with the Babcock - Leighton dynamo model. Foukal (1990) described the Babcock - Leighton dynamo model, through five stages of global magnetic field evolution, which is also concisely presented in Mulec (2007).

2.5 Meridional motions

Meridional motions are large-scale, longitudinally averaged, i.e., axisymmetric, north– south and radial velocity components on the Sun (Stix 1989). They lie in the meridional planes:

vm = (⟨vr⟩, ⟨vθ⟩, 0) (2.2) where (r, θ, ϕ) are the polar spherical coordinates. Only the latitudinal component vθ = vmer can be measured by tracers on the solar surface: ∂b v = (2.3) mer ∂t where b is the heliographic latitude. One important feature of the solar rotation profile is that lower latitudes rotate faster than higher latitudes which implies that most of the angular momentum of the Sun is located in the zones around the equator. Therefore, it is necessary to assume that there must be some mechanism which transfers angular momentum towards lower latitudes. The model of axisymmetric meridional circulation (Sec. 3.2), is dealing only with the meridional motions shown in the left upper part of the Fig. 3.1, neglecting any other effect that could produce the angular momentum transport (Gilman 1981; Stix 1989). This model could produce transfer of the angular momentum towards the equator, but it seems unreliable, as the model requires surface meridional motions in the direction towards the solar equator, which contradicts the observations in many published obser- vational results listed below (Paper V - Sec. 4.5). According to references listed below, there is a huge variety of possible orientations of the meridional flow:

• outwards from the centre of activity (equatorwards meridional motions at low latitudes and polewards meridional motions at medium latitudes)

– sunspot groups as tracers (Tuominen 1961; Tuominen & Kyrolainen 1982; Tuominen & Virtanen 1984; Balthasar et al. 1986; Howard 1991; Kambry et al. 1991; Lustig & Wöhl 1994; Wöhl & Brajša 2001) – sunspots and sunspot groups as tracers (Howard & Gilman 1986)

19 • towards the centre of activity (opposite orientation than in the previous item)

– solar plages as tracers (Howard 1991) – sunspot groups as tracers – Sudar et al. (2014), Sudar et al. (2017) - Paper V (using the arguments from Olemskoy & Kitchatinov (2005), explained below)

• different behaviour around the centre of solar activity

– small magnetic features analysed with different methods (Komm et al. 1993; Snodgrass & Dailey 1996) – CBPs (Vršnak et al. 2003)

• poleward meridional flow

– Doppler measurements (Duvall 1979; Howard 1979; Labonte & Howard 1982; Snodgrass 1984; Hathaway 1996) – CBPs (Sudar et al. 2016) - Paper IV – high - resolution magnetograms (Komm et al. 1993) – time - distance helioseismology (Zhao & Kosovichev 2004; González Hernández et al. 2008, 2010; Kosovichev & Zhao 2016)

• equatorward meridional flow

– Doppler measurements (Perez Garde et al. 1981).

In contrast to other Doppler measurements, Perez Garde et al. (1981) show merid- ional motions in the direction towards the solar equator. Sometimes, using the Doppler data, almost no meridional flow is detected (Lustig & Wöhl 1990). Doppler measure- ments may be affected by the center-to-limb variations of turbulent broadening and asymmetry of spectral lines (Kambry et al. 1991). In Sec. 4 of Paper V (Sec. 4.5), we attempt to explain the wide variety of results for meridional flow in a consistent way. We suggest that the difference between the observed meridional motion of sunspot groups and Doppler measurements, as well as the difference between sunspot groups and some other tracer measurements (e.g. CBPs measurements), lies in the fact that sunspot groups are located within active regions where the meridional motion is different than outside of those areas. In Paper V, the different observations of the meridional motion of sunspot groups, towards the centre of activity or oppositely, are suggested to be a consequence of the

20 assumption that it is irrelevant to which latitude one assigns the observed meridional velocity. Olemskoy & Kitchatinov (2005) have pointed out that in order not to detect false meridional flows due to uneven distribution of tracers across different latitudes, it is necessary to assign calculated velocities to the latitude of the first measurements of position for each tracer.

2.6 Horizontal Reynolds stress

The transport of angular momentum towards the solar equator is necessary for explain- ing the observed solar differential rotation profile. The origin of the differential rotation is theoretically reviewed in Rüdiger (1980), who defines Reynolds stress, meridional motions and magnetic field as possible sources of angular momentum transport. Brun & Rempel (2009) showed that the actual solar rotation profile Ω(r, θ) really is the result of the mutual contribution of meridional circulations, Reynolds stress transports and magnetic field. More details on the theory of the Reynolds stresses and angular momen- tum transport are available in Sections 3.3 and 3.4. The existence of angular momentum transport has been observationally revealed by Ward (1965), where the Reynolds stress was considered as the main contributing factor. From Ward (1965) onward, many obser- vational results related to this topic have been published, e.g. Pulkkinen & Tuominen (1998); Vršnak et al. (2003); Rüdiger & Hollerbach (2004); Sudar et al. (2014). It is clear that very often the transport of angular mometum is attributed primarly to horizon- tal Reynolds stress, which represents an interdependence between meridional motions vmer and rotation velocity residuals Δvrot. Horizontal Reynolds stress is a particularly important parameter which gives us the information whether the rotation velocity residuals and meridional velocities are related. It actually represents the covariance Q of the two velocities, one in the θ and another in ϕ direction (eq. 3.9)

Q = ⟨vmerΔvrot⟩ (2.4) where the brackets represent averaging over the longitude, vmer meridional velocity (eq. 2.3) and Δvrot rotation velocity residuals (Schröter 1985):

vrot = ωr⊙ cos b (2.5)

Δvrot = vrot − ⟨vrot⟩ (2.6) The covariance Q is the horizontal component of the Reynolds stress tensor (eq. 3.6) which might generate an equatorward transport of angular momentum (Ward 1965; Schröter 1985; Rüdiger 1989; Pulkkinen & Tuominen 1998; Stix 2002). The negative

21 Figure 2.9: Relationship between covariance, Q, and latitude, b, represented by average values in bins of different width obtained with data from Sudar et al. (2014) (”Sudar”), Ward (1965) (”Ward”) and the results from Vršnak et al. (2003) (”Vršnak 10” and ”Vršnak 30”, for their bins of 10 and 30, respectively). The linear fit derived by Sudar et al. (2014) is shown with a solid thin line. sign of Q represents the solar-like differential rotation with the equator rotating faster than the poles (Rüdiger 1989). A relationship between the meridional velocities and rotation velocity residuals is very often investigated, by plotting them in the vmer − Δvrot parameter space and calcu- lating the best linear fit values, in order to reveal whether these velocities are mutually dependent or independent (Paper V - Figure 4). If the slope of the linear fit of the vmer − Δvrot parameter space is negative, it tells us that on average the angular momen- tum is indeed transferred from higher to lower latitudes. Velocity components with just a few m/s are able to generate horizontal Reynolds stress of the order of several 103 m2/s2, which is sufficient to maintain the observed solar rotation profile (Schröter 1985). Several 103 m2/s2 for horizontal Reynolds stress were obtained by different authors, confirming the existence of the transport of the angular momentum towards the equator (Ward 1965; Belvedere et al. 1976; Schröter & Wöhl 1976; Gilman & Howard 1984a; Pulkkinen & Tuominen 1998; Vršnak et al. 2003; Sudar et al. 2014). A latitudinal dependence of the relative horizontal Reynolds stress was investigated and an increase with the latitude was found in some studies, with an increase of up to about 30◦ of latitude, where there is a possible maximum (Ward 1965; Gilman & Howard 1984a; D’Silva & Howard 1995; Meunier et al. 1997; Pulkkinen & Tuominen 1998; Vršnak et al. 2003; Sudar et al. 2014). Some of these results are presented in

22 Figure 2.10: Distribution of the horizontal Reynolds stress in the new model of the angular momentum trans- port, model M3, from Pipin & Kosovichev (2018).

Fig. 2.9, taken from Sudar et al. (2014). This phenomenon is also clearly visible in Fig. 2.10, from the distribution of the horizontal Reynoldss stress derived by a new model of the angular momentum transport, model M3 (Pipin & Kosovichev 2018). The model M3 shows a positive horizontal Reynolds stress in the middle of the convection zone of the northern hemisphere (red), while negative values are derived for the bot- tom of the convection zone and in the subsurface shear layer (blue). The comparison of the model results with the measurements presented in Fig. 2.9, where all results are derived from sunspot groups except Vršnak results which originate from CBP-s mea- surements, yields the conclusion that these features (sunspot groups, CBPs) are likely rooted just below the subsurface shear layer. The same conclusion has been derived by Pipin & Kosovichev (2018), by comparing the model results with the measurements from sunspot motions presented in Paper V (Figure 5).

2.7 Relationship between the solar rotation and activity

Numerical models of the interactions of convection and rotation with the magnetic field in deep spherical shells showed that an increase in magnetic field would result in the reduction of differential rotation (Brun et al. 2004). This reduction of differential rotation is due to the braking effect exerted by non-axisymmetric magnetic fields (subsections 3.1 and 3.2). Lanza (2006, 2007) analytically solved the angular momentum transport equation within the convection zone of a rotating star and obtained a similar result. This means that in the presence of strong magnetic fields, the net transport of angular momentum towards the equator is less efficient. In other words, at the minimum of activity, when the magnetic field is weaker, the angular momentum is transported more efficiently towards the equator, which results in the observed: a) increased equatorial rotation reflected in larger differential rotation parameter A and b)

23 more pronounced differential rotation at the minimum of magnetic activity, reflected in larger absolute value of parameter B. Conclusion a) is in accordance with the experimental results. The higher average rotation velocities in the minimum compared to other phases of cycle, i.e. the inverse correlation between solar equatorial rotation rate (differential rotation parameter A) and activity, were found by many authors using different methods and data (Lustig 1983; Gilman & Howard 1984b; Balthasar et al. 1986; Gupta et al. 1999; Zuccarello & Zappalá 2003; Brajša et al. 2006; Jurdana-Šepić et al. 2011; Li et al. 2014; Badalyan & Obridko 2017). In Jurdana-Šepić et al. (2011) - Paper II, using rotation rates obtained by tracing CBP, the relationship between the solar rotation parameters and solar activity that is expressed by the sunspot number was investigated and a significant negative correlation for parameter A was found. Li et al. (2014) revisited published solar rotation data obtained from Hα charts and sunspots and confirmed the secular deceleration and negative correlation between the differential rotation parameter A and solar activity. Most recently, Badalyan & Obridko (2017) analysed the coronal green-line data from 1943 up to the present and found that the equatorial rotation rate increases in the epochs of minimum between the even- and odd-numbered solar cycles, and it reaches its minimum between the odd and even cycles. Conclusion b) actually means that the Sun rotates more rigidly at the maximum (a lower absolute value of the differential gradient) than at the minimum of activity (Lanza 2006, 2007; Brun et al. 2004). It is important to note that the values of B are negative, so a lower absolute value of B during the maximum of activity indicates a positive correlation, while a higher absolute value of B indicates a negative correlation. This means that the positive correlation obtained by theoretical predictions is in contrast to the observations, where very often a negative correlation is obtained. For instance, Jurdana-Šepić et al. (2011) - Paper II showed a small, statistically insignificant, negative correlation for the differential rotation parameter B. Badalyan & Obridko (2017) showed that the differential gradient increases in the even cycles, with its highest absolute values near the maximum of the activity cycle. To gain insight into these discrepancies between observations and theoretical predictions, in the Paper III (Ruždjak et al. 2017) the relationship between solar activity and differential rotation parameters was determined from the sunspot group position data from three data sets (GPR, USAF/NOAA and DPD). The results, although not statistically significant for all datasets, show a negative correlation between the solar equatorial rotation rate and solar activity, while the results for the differential rotation parameter B are different for the various datasets. The GPR data (1874 – 1976) result in an insignificant negative trend, while the data from USAF/NOAA and DPD (1977 – 2016) result in a positive correlation with a 2σ statistical significance. The

24 results could imply that the correlation of the differential rotation parameter B has changed (period 1977 – 2016, positive correlation; period 1874 – 1976, insignificant negative correlation). When the sets are combined, the larger GPR dataset prevails and negative correlation is observed. It is obvious that further analysis are necessary, so a continuation of the analysis of the KSO data presented in the Paper I is in preparation. In this analysis, the temporal variation of the solar rotation parameters (cycle to cycle variations, variations in single years, variations during the solar cycle) and the relationship between the parameters of the solar rotation and the solar activity, using the yearly values of the solar rotation parameters, has been investigated. One of the conclusions made so far is that more rigid rotation, i.e., a lower absolute value of the differential gradient, has been observed during the minimum of activity, which is in contradiction with theoretical assumptions. The KSO data has also never been used before for this purpose for the years after 1980, as well as for an analysis of meridional motions and horizontal Reynolds stress (Ruždjak et al. (2018) - Paper VIII).

25 3 Theory of the solar differential rotation

Since the discovery of differential rotation of the Sun by Carrington (1863) until the emergence of helioseismology, different concepts of the differential rotation theory were formulated. Helioseismology helped us to improve our understandings of the internal solar rotation, as well as enabled us to check if numerical models based on the theory reproduce solar rotation quite well or not. As pointed out in Tassoul & Tassoul (2004), the existing concepts turned out to be in a good accordance with observed rotational and dynamo effects and almost all of them are dealing with the transport of angular momentum inside the Sun. Namely, the angular momentum transport seems to be the cause of occurrence of differential rotation (Foukal 1990; Stix 2002).

3.1 Angular momentum transport and magnetic braking

It is confirmed by several observational and theoretical results that the Sun was a very fast rotator during its initial state, where by „initial” the beginning of the hydrogen- burning phase, e.g. the Sun’s main-sequence life is assumed (Stix 2002). As shown by Schatzman (1959), the magnetic braking is responsible for the loss of the angular momentum. Pinsonneault et al. (1989) assumes that the total initial angular momentum 42 43 2 J0 is between 5 × 10 and 5 × 10 kgm /s, while Rüdiger & Kitchatinov (1996) takes into account a rotation rate several tens of times larger than the present. The effect of magnetic braking assumes the angular momentum reduction, caused by escaping of the matter from the surface of the Sun that takes with it some angular momentum. So, the whole angular momentum of the Sun lowers and with it the solar rotation velocities. The total loss of the angular momentum, e.g. the total torque exerted upon the Sun by the magnetic braking, is presented as the time derivative of the angular

26 momentum, J˙ :

˙ ≃ 2 J ΩrAm˙ (3.1) where Ω is angular velocity which multiplied with rA gives linear (azimuthal) veloc- * ity, rA is Alfv`en radius which represents the lever arm for the torque and m˙ is a time derivative of mass, e.g. the mass loss (Stix 2002).

3.2 Axisymmetric and non-axisymmetric convective motions

Plasma circulations in the solar convective zone also carry with their mass angular mo- mentum, so any flow that has a component in either a radial or latitudinal direction can change the rotation profile in the meridional plane. Therefore, these circulations have been described until now by two different transporting mechanisms of the angular mo- mentum (Gilman 1981): axisymmetric meridional circulations and non-axisymmetric circulations represented by the velocity correlations called Reynolds stresses (Fig. 3.1). These two processes both contribute to the angular momentum transport in a rotating, convective shell. Taking into account all the scales of motion, it is possible to derive radial and latitudinal angular velocity profiles (Foukal 2013). In the upper part of the Fig. 3.1, two meridional circulations of opposite orientation are shown. We will focus on the left one, as it may produce a net transport of angular momentum towards the equator resulting in equatorial acceleration which agrees with the observed angular velocity profiles. The left upper part shows the starting state which assumes nearly constant angular velocity (the solid body rotation) and the absence of any other angular momentum transport mechanism. Then, the fluid moving toward the equator on the outer branch and crossing the dashed line, would contain more angular momentum than fluid moving toward the pole on the inner branch, where the moment arm is shorter. This explains the transfer of the angular momentum from the pole to the equator (Gilman 1981; Foukal 2013). In the lower part of Fig. 3.1, angular momentum transport by non-axisymmetric motions is illustrated. Non-axisymmetric convective motions can be produced through correlations between east-west, north-south or radial motions, called Reynolds stresses. The left lower part of Fig. 3.1 shows a horizontal flow pattern that leads to angular momentum transport toward the equator. It consists of two flows: flow towards the di- rection of rotation that is inclined to the equator and opposite flow that is inclined toward

* The Alfv`en radius rA is the distance at which escaping material, which encompass open field lines originating from the flux tubes, rotate nearly rigidly. In other words, it is the distance where the magnetic energy density is equal to the kinetic energy density, which means that beyond rA the field becomes too weak to enforce rigid rotation.

27 Figure 3.1: Two main transporting mechanisms of the angular momentum: axisymmetric convective motions represented by meridional circulations (upper part) and non-axisymmetric convective motions represented by the correlations called Reynolds stresses (lower part). Taken from Gilman (1981). the pole. If the average value of the cell velocities is determined over the longitudes, e.g. along the dashed line, resulting orientation is equatorward, which as a consequence implies equatorward angular momentum transfer (Gilman 1981; Foukal 2013). In the lower right part of Fig. 3.1, a typical convective circulation pattern in a local longitude radius plane is shown. In this case the net angular momentum flux is oriented inward (Gilman 1981).

3.3 The Λ-effect and eddy viscosity

As we consider the solar convective zone as a spherical shell with turbulent convection, convective velocity can be described by the velocity field in spherical symmetry

u = (ur, uθ, uϕ) (3.2) Motions of the convective flows are disturbed by the two main forces (Gilman 1981): the Coriolis force due to solar rotation that acts in horizontal direction and the buoyancy force that acts in radial direction. So, as shown by Kitchatinov (2005) and then by Kitchatinov (2011), the convective turbulent mixing produces two different

28 Figure 3.2: Two different cases of angular momentum transport produced by the convective turbulent mixing (from Kitchatinov (2005)). Arrows on the top show the direction of rotation. Dashed arrows show original motions; solid arrows show disturbed motions. Left part: disturbings caused by the Coriolis force, downward transport of angular momentum. Right part: disturbings caused by the buoyancy force, upward transport of angular momentum. cases of angular momentum transport (Fig. 3.2). Left part of the Fig. 3.2 shows the original radial motion (dashed arrows) deflected by the Coriolis force in horizontal direction (solid arrows), causing the formation of azimuthal velocities and transporting the angular momentum downward, into the depths of the convection zone. Right part of the Fig. 3.2 shows the original horizontal motion deflected by the buoyancy force. In this case, the angular momentum is transported upward by the horizontal convective mixing. Observing of these cases in particular does not show a complete picture. The net flux of the angular momentum is a result of the anisotropic mixing of averaged radial and horizontal motions with different intensities (Kitchatinov 2005, 2011):

Λ ∗ ⟨ 2⟩ − ⟨ 2⟩ Qrϕ = Ω ( uϕ ur ) sin θ (3.3) where Qrϕ represents the cross-component of the correlation tensor Qij (eq. 3.6), described in more detail below, while upper index Λ indicates the Λ-effect. Here, the Coriolis number Ω∗ = 2τΩ (3.4) where τ is a convective turnover time and Ω angular velocity, represents a key parameter of the differential rotation theory and gives an information about the intensity of interaction between the turbulent convection and rotation (Kitchatinov 2005, 2011). Λ ̸ Eq. 3.3 shows non-vanishing Reynolds stress in the radial-azimuthal plane, Qrϕ = 0. This means that radial and azimuthal motions in the original turbulence are different, therefore radial angular momentum flux arises. When the differences between the azimuthal and meridional motions are present, e.g. non-vanishing Reynoldss stress in Λ ̸ the horizontal plane is present, Qθϕ = 0, a meridional angular momentum flux arises (Kitchatinov 2005).

29 The Λ-effect implies appearance of the turbulent fluxes of angular momentum in the case of rigid (uniform) rotation (Rüdiger 1980, 1989). The Λ-effect is considered to be the basic source of differential rotation, firstly noted by Lebedinskii (1941), and then investigated by Biermann (1951); Kippenhahn (1963); Köhler (1970). Kitchatinov & Rüdiger (1993) were first to provide the original theoretical formulation, with an improved representation in Kitchatinov & Rüdiger (2005), followed by a series of theoretical considerations of which one of the most recent is Pipin & Kosovichev (2018). In addition to Λ-effect there is another effect, the eddy viscosity, which represents the most widely known effect of turbulence on large-scale fields. Therefore, it is often called the turbulent or anisotropic viscosity. It has the ability to transport angular momentum also in the case when the rotation is not uniform. So, most differential rotation models introduce a decomposition of properties into a non-diffusive (Λ) and diffusive component, in order to understand the maintenance of large scale flows. Therefore, the total velocity is divided into a mean component V and fluctuating component u (eq. 3.2)

v = V + u, V = ⟨v⟩ (3.5) where fluctuating component (convective turbulent velocity) vanishes when averaged, ⟨u⟩ = 0. Generally, Reynolds stress tensor is defined as

Rij = −ρQij, Qij = ⟨uiuj⟩ (3.6) which actually describes the effect of turbulence on the mean, large-scale flow. Most differential rotation models divide the Reynolds stresses in their non-diffusive part (upper index Λ) and diffusive part (upper index υ), according to the previously mentioned decomposition,

Λ υ Qij = Qij + Qij (3.7) Λ where the first part on the right-hand side, Qij corresponds to the Λ-effect and the second υ called the viscous part, Qij, to the contribution from the eddy viscosity (Kitchatinov 2005, 2011). The radial and horizontal Reynolds stresses, that arise from eq. 3.6, are defined as

Qrϕ = ⟨uruϕ⟩ (3.8)

Qθϕ = ⟨uθuϕ⟩. (3.9)

30 Rüdiger (1980) provided expressions for the radial and horizontal Reynolds stresses, including non-diffusive (Λ-effect) part proportional to the angular velocity Ω and diffusive (viscous) part proportional to the gradient of Ω: ∂Ω Q = Λ sin θΩ − υ r sin θ (3.10) rϕ r t ∂r ∂Ω Q = Λ cos θΩ − ˆsυ sin θ (3.11) θϕ h t ∂θ where Λr and Λh are radial and horizontal parameters that describe the non-diffusive transport, ˆs is a dimensionless parameter which indicates that the diffusive effect can be of different strength in the vertical and horizontal direction and υt is the kinetic viscosity. Afterwards, Kitchatinov & Rüdiger (1993) (their eqs. 4.2 a, 4.2 b), Kitchatinov & Rüdiger (2005) (their eq. 3) and Pipin & Kosovichev (2018) (their eqs. 10 and 12) provided improvements of the expresions, by defining Λr = υtV and Λh = υtH, where V and H are the normalized vertical and horizontal fluxes. Rüdiger & Tuominen (1987) pointed out that Λh has to be non-zero, because otherwise in any model which yields the correct ∂Ω/∂θ, Qθϕ contradicts the observed sign (see equations 3.10 and 3.11). According to eq. 3.9, in Sec. 2.6 horizontal component of the Reynolds stress tensor, used in the thesis (Paper IV and V), was defined (eq. 2.4). Observations yield the full correlation represented by eq. 3.9, which includes contributions from both the Λ- effect and eddy viscosities. Model calculations by Küker et al. (1993) yields that sign of the horizontal Reynolds stress on the solar surface has to be negative. This is in accordance with the observational results, e.g. Nesme-Ribes et al. (1993); Sudar et al. (2014); Ruždjak et al. (2018); Paper IV and Paper V included in the thesis.

3.4 Angular momentum transport

To produce differential rotation, the turbulent convective motions must transfer angular momentum. So, the theory of nonuniform rotation is very closely connected with the theory of angular momentum conservation (Rüdiger & Hollerbach 2004). Angular momentum transport equation can be inferred from the equation of motion as shown in Stix (1989, 2002), as well as in Kitchatinov (2005, 2011). This considerations omit the influence of the magnetic field on the redistribution of angular momentum, and focus only on the meridional motions and Reynolds stresses. In Rüdiger & Hollerbach (2004) and Brun & Rempel (2009) angular momentum transport equation is modified by inclusion of the magnetic fields, e.g. more than two terms (meridional flow and Reynolds stresses) now enter the balance:

31 { ( )} ∂Ω r sin θ ρr2 sin2 θ + ▽ ρr2 sin2 θΩv + ρr sin θ⟨u u⟩ − ⟨B⟩⟨B ⟩ + ⟨B′B′ ⟩ = 0 ∂t m ϕ 4π ϕ ϕ (3.12) where ρ is the mass density, Ω is angular velocity, convective velocity u is defined by eq. 3.2, ⟨uϕu⟩ is equal to (Qrϕ, Qθϕ, 0) (see eqs. 3.8 and 3.9) and the magnetic field is decomposed into mean and fluctuating part

B = ⟨B⟩ + B′. (3.13) Divergence in eq. 3.12 is applied on the vector that represents angular momentum flux, i.e. this expression represents the conservation of angular momentum. So, the transport of the angular momentum can be described by meridional circulation vm (first term in divergence of eq. 3.12), Reynolds stresses (second term in divergence of eq. 3.12) and Maxwell stresses (third term in divergence of eq. 3.12). Thereby, the actual solar rotation profile Ω(r, θ) can be represented as a result of the mutual contribution of the meridional circulations, Reynolds stress transports and Maxwell stresses (Rüdiger & Hollerbach 2004; Brun & Rempel 2009; Foukal 2013). If we assume rigidity of the rotation at some moment and if the divergence from the eq. 3.12 is non-zero, the rotation becomes non-uniform. In the regions where the divergence of the angular momentum flux is positive, deceleration of the rotation is present, and vice versa, where flux is negative, acceleration is present. According to the decomposition of the magnetic field (eq. 3.13), the Maxwell stresses are separated into the contributions from the mean field and fluctuating field. Maxwell stresses actually represent a backreaction (through the Lorentz forces) on large scale flows, induced with the magnetic field produced by the solar dynamo. They actually oppose the Reynolds stresses, contribute to the transport of the angular momentum towards the solar poles and consequently affect the reduction of the differential rotation (Brun et al. 2004). So, a more pronounced differential rotation, i.e. a higher equatorial rotation velocity (parameter A) is expected when magnetic fields are weaker, as shown in Paper II and III.

3.5 Models of the differential rotation

All models of differential rotation calculate some motions explicitly, and parametrize the effects of the others. In mean field models, only the meridional motions and differential rotation are explicitly determined, and all other motions are parametrized. Another class of models calculates global-scale convection, including departures from axisymmetry, and parametrizes smaller scale motions (Gilman 1981; Rüdiger 1989;

32 Rüdiger & Hollerbach 2004). In general, mean field models and local numerical simulations show a qualitative agreement, while larger differences occur when the entire convection zone is observed. According to the high parametrization, mean-field models are not highly accurate, but are more widely used. Numerical calculations for a small part of the convection zone would be able to test the mean-field theory (Rüdiger et al. 2005). A brief history of development of the models is given in Kitchatinov (2005).

3.5.1 Mean-field models A mean-field theory of stellar convection and rotation has been developed by Rüdiger (1989), while improvements of the theory are given in Kitchatinov & Rüdiger (1993); Rüdiger & Hollerbach (2004); Kitchatinov & Rüdiger (2005); Pipin & Kosovichev (2018). Fig. 3.3 (left side) shows the distribution of the rotation rate in the solar convective zone and streamlines of the meridional circulation (right side) for one of the most recently developed mean-field models (Pipin & Kosovichev 2018). The mean field models are based on eq. 3.12, i.e. on decomposition of properties into their large scale mean values (differential rotation, meridional flow, large scale magnetic field) and small scale fluctuating parts (Reynolds stresses, small scale magnetic field). The transport of the angular momentum depends on two types of velocity structures: meridional motions and Reynolds stresses. Both types of transport occur simultaneously, but mean-field models consider the Reynolds stresses as the more fundamental ones. There are two classes of the mean-field models of the differential rotation: „anisotropic viscosity” models and „latitude-dependent heat transport” models. But, the construction of realistic models, which satisfy observational constraints, requires the inclusion of both the Λ-effect and eddy viscosity contribution (Sec. 3.3), as well as thermodynamic contributions (Pidatella et al. 1986). The slightly hotter polar regions in comparison to the equatorial zone are produced by the influence of rotation on the convective heat transport. The pole-equator temperature difference was measured by Rast et al. (2008), suggesting that the solar poles are warmer for about 2.5 K. As pointed out in Foukal (2013), the temperature difference is of a very small scale but this energy balance within the convection zone seems to be an angular momentum transport driver. Weiss (1965) and Durney & Roxburgh (1971) assumed that convective heat flux FC depend on latitude. An expression for the convective heat flux is ∂ C − S Fi = ρTχij (3.14) ∂rj where ρ is density, T temperature, S the specific entropy and χij the thermal diffusivity tensor (Kitchatinov et al. 1994; Kitchatinov 2011).

33 Figure 3.3: Angular velocity pattern Ω(r, θ)/2π in the solar convection zone (a) and meridional circulation pattern (b) developed by mean-field model M3 in Pipin & Kosovichev (2018)

The profile of differential rotation cannot be explained by considering only angular momentum equation (eq. 3.12). Thermal effects included in eq. 3.14 should not be neglected, as well as an equation that describes the origin of the meridional flows:

∂Ω2 g ∂S D(vm) = r sin θ − (3.15) ∂z cpr ∂θ where z = r cos θ is the distance from the equatorial plane, S is the specific entropy, g is gravity, and cp is the specific heat at constant pressure. Difference of the two terms on the right hand side represent the main cause of the meridional flow. The first term is consequence of the Coriolis force, while the second term is consequence of the pressure and buoyancy forces (Brun & Rempel 2009; Kitchatinov 2011). For our discussion, there is no need to show explicitly left part of the equation. As it is pointed out in Brun & Rempel (2009) and Kitchatinov (2011), the angular momentum equation (eq. 3.12) together with the meridional flow equation (eq. 3.15) and entropy equation with the heat flux (eq. 3.14) represent a set of equations that enables numerical solution for the distributions of angular velocity, meridional flow and entropy in a stellar convection zone. In this way free parameters used in the mean-field models are reduced and uncertainty of the mean-field models is minimized.

3.5.2 Global (explicit) models Mean-field models determine the global flows by averaging equations of turbulent flows. Global models represent another approach which consists of solving together all the hydrodynamic equations in spherical geometry, by numerical calculation, while averaging is applied afterwards, on obtained solutions (Kitchatinov 2005).

34 First calculations of the global convection models are presented in Gilman (1983), Glatzmaier (1985, 1987), Gilman & Miller (1986). Miesch et al. (2000) simulated the global convection by solving the time-dependent equations of motion. These results are in contrast to the helioseismological results from Libbrecht (1988), which show a nearly conical isorotation surfaces. As pointed out in Stix (1989, 2002), these models yield the correct surface rotation law, axisymmetric meridional circulation and differences in surface temperatures, as well as correct horizontal Reynolds stresses, Qθϕ. Recent observations, concerning the global modelling, are trying to provide joint models of the differential rotation and the solar dynamo, in the same time dealing with the plasma motions and the magnetic field and using different numerical techniques (Guerrero et al. 2013; Kitchatinov & Nepomnyashchikh 2016).

35 4 Peer-reviewed publications

During my PhD studies I contributed to 13 research papers, 8 of which are peer-reviewed publications indexed in the Current Contents/Web of Science, not all of them covering the field of solar physics. Five peer-reviewed publications are within the thesis topic and are presented hereafter (Sec. 1, Paper I - Paper V from the List of Publications).

4.1 Paper I: Solar differential rotation in the period 1964 - 2016 de- termined by the Kanzelhöhe data set

The following article is published in Astronomy and Astrophysics, Volume 606, Issue A72, pp. 1-10 (2017). It is reproduced by the permission of ESO. The .pdf document is available in the online version of this journal and hereafter. My own contribution to this work was 70%.

The main aims of this paper are presented in Sec. 1. The main focus of this paper is to analyse the solar differential rotation by tracing sunspot groups during the 1964 - 2016 period, using the KSO sunspot drawings and white light images, which have been used before for this purpose just until 1980’s. The main conclusion of this paper is that KSO provides a valuable data set with a satisfactory accuracy. The comparison of calculated rotation profiles derived from the KSO data in this work and from all other studies collected from different sources yielded a conclusion about the KSO accuracy after 1960’s. The KSO data used in this paper, for the time period from 1964 till present, is of a comparable accuracy with the DPD or GPR data sets. Although there were some systematic errors present in the past, this data set is well suited for long term studies and therefore still continued. Also, it is

36 shown that the quality of the KSO sunspot drawings has gradually increased during the last 50 years.

37 A&A 606, A72 (2017) Astronomy DOI: 10.1051/0004-6361/201731047 & c ESO 2017 Astrophysics

Solar differential rotation in the period 1964–2016 determined by the Kanzelhöhe data set

I. Poljanciˇ c´ Beljan1, R. Jurdana-Šepic´1, R. Brajša2, D. Sudar2, D. Ruždjak2, D. Hržina3, W. Pötzi4, A. Hanslmeier5, A. Veronig4, 5, I. Skokic´6, and H. Wöhl7

1 Physics Department, University of Rijeka, Radmile Matejciˇ c´ 2, 51000 Rijeka, Croatia e-mail: [email protected] 2 Hvar Observatory, Faculty of Geodesy, University of Zagreb, Kaciˇ ceva´ 26, 10000 Zagreb, Croatia 3 Zagreb Astronomical Observatory, Opatickaˇ 22, 10000 Zagreb, Croatia 4 Kanzelhöhe Observatory for Solar and Environmental Research, University of Graz, Kanzelhöhe 19, 9521 Treffen am Ossiacher See, Austria 5 Institute for Geophysics, Astrophysics and Meteorology, Institute of Physics, University of Graz, Universitätsplatz 5, 8010 Graz, Austria 6 Astronomical Institute of the Czech Academy of Sciences, Fricovaˇ 298, 25165 Ondrejov,ˇ Czech Republic 7 Kiepenheuer-Institut für Sonnenphysik, Schöneckstr. 6, 79104 Freiburg, Germany

Received 26 April 2017 / Accepted 5 July 2017

ABSTRACT

Context. Kanzelhöhe Observatory for Solar and Environmental Research (KSO) provides daily multispectral synoptic observations of the Sun using several telescopes. In this work we made use of sunspot drawings and full disk white light CCD images. Aims. The main aim of this work is to determine the solar differential rotation by tracing sunspot groups during the period 1964–2016, using the KSO sunspot drawings and white light images. We also compare the differential rotation parameters derived in this paper from the KSO with those collected fromf other data sets and present an investigation of the north – south rotational asymmetry. Methods. Two procedures for the determination of the heliographic positions were applied: an interactive procedure on the KSO sunspot drawings (1964–2008, solar cycles Nos. 20–23) and an automatic procedure on the KSO white light images (2009–2016, solar cycle No. 24). For the determination of the synodic angular rotation velocities two different methods have been used: a daily shift (DS) method and a robust linear least-squares fit (rLSQ) method. Afterwards, the rotation velocities had to be converted from synodic to sidereal, which were then used in the least-squares fitting for the solar differential rotation law. A comparison of the interactive and automatic procedures was performed for the year 2014. Results. The interactive procedure of position determination is fairly accurate but time consuming. In the case of the much faster automatic procedure for position determination, we found the rLSQ method for calculating rotational velocities to be more reliable than the DS method. For the test data from 2014, the rLSQ method gives a relative standard error for the differential rotation parameter B that is three times smaller than the corresponding relative standard error derived for the DS method. The best fit solar differential rotation profile for the whole time period is ω(b) = (14.47 ± 0.01) − (2.66 ± 0.10) sin2 b (deg/day) for the DS method and ω(b) = (14.50 ± 0.01) − (2.87 ± 0.12) sin2 b (deg/day) for the rLSQ method. A barely noticeable north – south asymmetry is observed for the whole time period 1964–2016 in the present paper. Rotation profiles, using different data sets, presented by other authors for the same time periods and the same tracer types, are in good agreement with our results. Conclusions. The KSO data set used in this paper is in good agreement with the Debrecen Photoheliographic Data and Greenwich Photoheliographic Results and is suitable for the investigation of the long-term variabilities in the solar rotation profile. Also, the quality of the KSO sunspot drawings has gradually increased during the last 50 yr. Key words. Sun: photosphere – Sun: rotation – sunspots

1. Introduction (EGR) data set (Pulkkinen & Tuominen 1998; Javaraiah 2003; Zuccarello & Zappalá 2003; Javaraiah et al. 2005; Javaraiah & Solar rotation can be generally determined with three meth- Ulrich 2006; Brajša et al. 2006, 2007; Sudar et al. 2014); the ods: the tracer method, the spectroscopic method, and the Kodaikanal data set (Gupta et al. 1999); the Debrecen photohe- helioseismology method (Stix 2004). Sunspots and sunspot liographic data (DPD; Sudar et al. 2017); and the Kanzelhöhe groups represent one of the most commonly used tracers in Observatory for Solar and Environmental Research (KSO) the literature. The main reason is the availability of long- data set (Lustig 1982, 1983; Balthasar & Fangmeier 1988; term data sets from various observatories: the Greenwich Poljanciˇ c´ Beljan et al. 2014; Pötzi et al. 2016). photoheliographic results (GPR) data set (Balthasar & Wöhl The KSO provides daily multispectral synoptic observations 1980; Arévalo et al. 1982; Balthasar et al. 1986b); the Mt. of the Sun using several telescopes (Veronig & Pötzi 2016). Wilson data set (Howard 1984; Gilman & Howard 1984; However, especially the sunspot drawings and white light im- Hathaway & Wilson 1990); the extended Greenwich results ages seem to be neglected by the scientific community. Previous

Article published by EDP Sciences A72, page 1 of 10 A&A 606, A72 (2017) studies performed on solar drawings from the KSO, in which sunpots and sunspot groups were analysed, involve data obtained through the year 1985 (Lustig 1982, 1983; Lustig & Dvorak 1984; Balthasar et al. 1986a; Hanslmeier & Lustig 1986; Lustig & Hanslmeier 1987; Balthasar & Fangmeier 1988; Lustig & Wöhl 1991). There are a few papers that use more recent sunspot drawings (Temmer et al. 2006; Poljanciˇ c´ et al. 2010; Poljanciˇ c´ et al. 2011; Poljanciˇ c´ Beljan et al. 2014). Temmer et al. (2006) provide the catalog of the hemispheric sunspot numbers and investigate the north – south activity asymmetries, but do not deal with the solar rotation. Poljanciˇ c´ et al.(2010) checked the precision of the Solar Optical Observing Network/United States Air Force/National Oceanic and Atmospheric Administration (SOON/USAF/ NOAA) data set by comparing it with the GPR data set. The KSO data set (heliographic coordinates determined from the sunspot drawings for the years 1972 and 1993) were used as Fig. 1. Projection device for making the draft of sunspot drawings. The a reference for the comparison of sunspot position measure- mirrors in the light path of the refractor have reversed the drawing, i.e., west is left. The enlargement of the solar disk to a diameter of about ments. The SOON/USAF/NOAA data were found to be some- 25 cm corresponds to a magnification factor of about 100. what less accurate than the GPR data. Poljanciˇ c´ et al.(2011) expanded that analysis on several more data sets, again by us- ing the KSO data set as a reference, for the comparison of sunspot position measurements. In almost all cases the latitude, longitude, and synodic angular velocity differences between the KSO and other data sets are the smallest for DPD, GPR, and SOON/USAF/NOAA. Since the GPR and SOON/USAF/NOAA data sets have already been widely used for the analysis of dif- ferential rotation of sunspot groups and its temporal variation, we decided that the next step will be to process the DPD and KSO data sets. An analysis concerning the KSO data set is pre- sented in this work, while an analysis of the DPD data set is made in Sudar et al.(2017). Poljan ciˇ c´ Beljan et al.(2014) cal- culated differential rotation parameters and showed differential rotation profiles for solar cycles Nos. 20 and 22 for the KSO data in a preliminary form. Our analysis is a continuation of the investigation done by Lustig(1983). We processed the KSO sunspot drawings for solar cycles Nos. 20–23 (1964–2008), and the KSO white light images for solar cycle No. 24 (2009–2016). For the first time, the whole solar cycle No. 21 is examined using the KSO data (previously, it was just partially processed), as well as solar cycles Nos. 22– 24. Solar cycle No. 20 is where our work and the work of Lustig (1983) overlap and is therefore suitable for comparison of the results. In this paper we study the solar differential rotation and we Fig. 2. Scanned white light image with added NOAA numbers and a make a comparison of the differential rotation parameters col- grid that shows P and B0 angles. The image is not rotated, but an inter- lected from different sources. In addition, we discuss the north- face for de-rotating the images into the east-west direction is available south asymmetry of the solar rotation profile. (2002-08-19 06:10:41 UT).

2. Instrumentation and measurements The white light images are made by a refractor (d/ f = 130/1950 mm) using a filter for 546 nm with a band pass of The KSO belongs to the University of Graz, Austria1. Sunspot 10 nm. During the period 1989–2007 white light images were drawings are made with a refractor (d/ f = 110/1650 mm) us- recorded on photographic film. Usually, three white light images ing the projection system that enlarges the disk image to 25 cm per day were produced. Until recently, these images were acces- in diameter (see Fig.1). The drawings are updated online al- sible only on transparent sheet films, but the scanning process is most every day, depending on weather conditions. In general, proceeding and they are now available online for the years after weather conditions are good and continuous operation yields 1992 (see Fig.2). The films are scanned with a photo scanner for about 300 days of observations per year. The KSO archive transparent film material and as an output FITS and JPEG files consists of more than 60 yr of sunspot drawings (Pötzi et al. are produced (Pötzi 2010). 2016). Otruba(2006) described the solar monitoring program In July 2007 the system was replaced by the KSO Pho- at Kanzelhöhe Observatory and the sunspot drawings in detail. tosphere Digital Camera (KPDC), Pulnix TM-4100CL (Pötzi 2010). To improve the image quality a 12 bit 2048 × 2048 pixel 1 https://www.kso.ac.at/index_en.php Pulnix RM-4200GE camera was installed in August 2015.

A72, page 2 of 10 I. Poljanciˇ c´ Beljan et al.: Solar differential rotation determined by the Kanzelhöhe data set

3. Determination of the heliographic positions From the fits header P and B0 were extracted, and from the corresponding data file solar radii and X and Y pixel coordinates Our investigation covers the time period 1964–April 2016. We were extracted. Using the calculations from Meeus(1991), we used two procedures to determine the heliographic positions of calculated heliographic coordinates of sunspot groups. sunspot groups: an interactive procedure, using KSO sunspot We also tried to apply the automatic method on white light drawings for the time period 1964–2008, and an automatic pro- images recorded on photographic film material (1989–2007), but cedure, using KSO white light images for the time period 2009– the results were not very reliable. The calculated equatorial ro- 2016. tation velocities ranged from less than 14 deg/day to more than In order to avoid solar limb effects leading to high posi- 15 deg/day. The reason for these uncertainties is the strong varia- tion uncertainties, we limited the data to ±58 deg in central tion of the density (blackening) due to the developing procedure meridian distance (CMD) which covers about 85% of the pro- of the film material. These variations lead to some very faint im- jected solar radius (Balthasar et al. 1986b). With this cutoff we ages where the solar limb is very difficult to detect, leading to obtained a sample of 12 152 sunspot groups that correspond to incorrect solar radii and as a consequence to incorrect sunspot some 70 000 individual positions of sunspot groups. Recurrent positions. sunspot groups are counted as many times as they appear.

4. Determining the rotation velocities 3.1. Interactive procedure (1964–2008, solar cycles Nos. 20–23) In order to calculate the synodic angular rotation velocities we used two methods: the daily-shift method (DS), where the syn- 2 A software package called Sungrabber (Hržina et al. 2007) that odic rotation velocities were calculated from the daily differ- determines the position of tracers is used for the interactive pro- ences of the CMD and the elapsed time t cedure. The area weighted centers of sunspot groups are esti- ∆CMD mated by the naked eye, thus giving more importance (by mov- ω = , (1) ing the center of gravity) to the umbra and penumbra that are syn ∆t more pronounced, i.e., occupy a larger area. and the robust linear least-squares fit method (rLSQ), where the In order to find out if different observers could affect the mea- synodic rotation velocities were calculated by fitting a line to surements, we performed a test: two observers independently de- the measured positions in time CMD(t) for each tracer. The syn- termined the positions of 10 single H- and J-type sunspot groups odic rotation velocity corresponds to the slope of the fit. Since (see Poljanciˇ c´ et al. 2010, Fig. 3). Differences between measure- the measured data occasionally have outliers, due to false iden- ments of the two observers are negligible (on average ≈0.1 deg, tification and other reasons, we used a robust fit, namely iter- always less then 0.2 deg) when compared with the differences atively reweighted least-squares with Huber’s t weighting func- between various observatories (≈0.5 deg). Based on these re- tion (Huber 1981). When applying the rLSQ method, the rotation sults all interactive coordinate determinations in the present velocities were calculated by fitting a line to at least three data work were done by two different observers using the Sungrab- points. If the number of position measurements, i.e., data points, ber software. is calculated for each sunspot group, the median value is 5, while Sunspot groups were identified with the help of the GPR the maximum value reaches 11. We note that even in the cases and DPD databases (Baranyi et al. 2016; Gyori˝ et al. 2017). The where only three data points is used by the rLSQ method, it is software Sungrabber offers the possibility of loading the GPR or still more than the two data points used when applying the DS DPD sunspot group heliographic positions, as well as Greenwich method. or NOAA/USAF sunspot group numbers, and mark them on the Angular rotation velocities were converted from synodic to sunspot drawing used. sidereal using the procedure described by Roša et al.(1995) and by Brajša et al.(2002a), which was improved by Skoki c´ et al. (2014). The calculated sidereal velocities, ω, were used in the 3.2. Automatic procedure (2009–2016, solar cycle No. 24) least-squares fitting to the solar differential rotation law Based on the algorithm by Watson & Fletcher(2011) sunspot ω(b) = A + B sin2 b, (2) groups and their properties (size, umbra, penumbra, position) are identified by morphological image processing of KSO white where b is the heliographic latitude, and A and B are the solar light images. This data is prepared every observing day by KSO differential rotation parameters. In both methods, DS and rLSQ, and available over KSO ftp server3 as raw fits images4 and we assigned the velocity to the latitude and time of the first data files5. Based on the assumption that images of better qual- measurement of position. When average latitudes are used, the ity show more detail, we selected for each single day the fits false meridional flows can be detected (Olemskoy & Kitchatinov file with the corresponding data file that contained the largest 2005), but for rotation analysis it probably represents a negligi- amount of information. ble effect. The data files provide the sunspot group center of gravity in pixel coordinates. Only the umbra pixels are used to calculate the center of gravity. For each sunspot group, a few separate um- 5. Results bra/penumbra pixel coordinates are also available, determined as 5.1. Comparison of automatic and interactive procedures the mean position of the umbra/penumbra pixels, which are only area weighted. A direct comparison of the interactive and the automatic proce- dure can only be presented for the year 2014, the only year for 2 http://www.zvjezdarnica.hr/sungrabber/sungrabb.html which both procedures were applied. This year was selected as it 3 http://cesar.kso.ac.at/main/ftp.php represents the maximum of solar cycle No. 24 after a deep mini- 4 ftp://ftp.kso.ac.at/phokaD/FITS/synoptic/ mum, when almost no sunspots and thus data points were avail- 5 ftp://ftp.kso.ac.at/sunspots/drawings/automatic/ able, and the first maximum covered by the KSO Photosphere

A72, page 3 of 10 A&A 606, A72 (2017)

Table 1. Results of fittings of the KSO data for the year 2014 to rota- method. This also means that 51% of the reduced calculated tional velocity profile Eq. (2). sidereal rotation velocities belong to the northern hemisphere, while 49% belong to the southern hemisphere (valid for both the Row Method ABNvl DS and the rLSQ). 1 int, DS 14.55 ± 0.06 −2.14 ± 0.83 989 In Table2 we present the results of fittings for solar cycles 2 aut, DS 14.42 ± 0.08 −0.98 ± 1.11 831 Nos. 20–24 separately (DS: rows 1–5 for N+S, rows 8–12 for N, rows 15–19 for S; rLSQ: rows 22–26 for N+S, rows 29–33 3 int, rLSQ 14.62 ± 0.08 −2.52 ± 1.11 218 for N, rows 36–40 for S); the results for the part of the data set 4 aut, rLSQ 14.64 ± 0.09 −3.40 ± 1.32 180 using the interactive procedure covering solar cycles Nos. 20– Notes. The differential rotation parameters are calculated by the inter- 23 (DS: rows 6, 13, 20; rLSQ: rows 27, 34, 41) and the re- active (int) and the automatic (aut) procedures and both hemispheres sults for the whole data set using the interactive and automatic together. All calculations are done for DS and rLSQ methods separately method together covering solar cycles Nos. 20–24 (DS: rows 7, and passed through ±58 deg CMD filter and 8–19 deg/day velocity 14, 21; rLSQ: rows 28, 35, 42). All results are calculated us- filter. The sidereal parameters and their standard errors are expressed ing both methods (DS and rLSQ), for both hemispheres together in deg/day. Nvl is the number of calculated rotation velocities. (N+S), and for the northern hemisphere (N) and the southern hemisphere (S) separately. The years specified in Table2 (time period) represent the starting and ending epochs of the corre- 0 10 20 30 40 50 sponding solar cycles. They are taken from Brajša et al.(2009), except the beginning of solar cycle No. 24, which was taken 14.6 14.6 from SILSO World Data Center(2015), Royal Observatory of 14.4 14.4 Belgium, Brussels6. 14.2 14.2 14.0 14.0 ) 5.3. Comparison with other data 13.8 13.8 In this section we present a comparison of the differential rota- 13.6 13.6 deg / da y tion parameters A and B collected from different sources, mostly / ( 13.4 13.4 ω those already mentioned in Sect.1 as interesting (GPR, EGR,

13.2 int, DS 13.2 DPD, KSO) dealing only with sunspots and sunspot groups as int, rLSQ 13.0 13.0 tracers (Table3). Table3 consists of three parts: the upper part au t, DS 12.8 aut, rLSQ 12.8 (rows 1–21) gives the results for both solar hemispheres together, the middle part (rows 22–32) gives the results for the northern 12.6 12.6 hemisphere, and the lower part (rows 33–43) gives the results 0 10 20 30 40 50 for the southern hemisphere. Figure4 only shows the di fferen- b / deg tial rotation profiles for the results derived from the KSO data Fig. 3. Differential rotation profiles for the year 2014 (see Table1). set (Table3, rows 17–21). Figure5 shows the di fferential rota- Sidereal rotation velocity is denoted by ω and heliographic latitude by b. tion profiles for several data sets and both hemispheres together (see corresponding rows in Table3), using only sunspot groups (Note (a)) or sunspots and sunspot groups (Note (c)) as trac- Digital Camera. We note that the years near the solar minimum ers. The differential rotation profiles derived from the KSO data usually lead to large statistical errors as the data set becomes set Lustig(1983; Fig.5, red line) and the Mt. Wilson data set very small, e.g., the year 2009 with 260 spotless days. Howard et al.(1984; Fig.5, pink dashed line) show significantly The differential rotation parameters A and B determined for lower values when compared to other results. More details are the year 2014 with both procedures and the velocities obtained given in Sect. 6.2. with DS and rLSQ methods for both hemispheres together are presented in Table 1. Corresponding differential rotation profiles are shown in Fig.3. More details are given in Sect. 6.1. 5.4. North–south asymmetry of the solar rotation The differential rotation parameters derived in the present pa- 5.2. Differential rotation from the KSO data per for the whole time period (1964–2016) and for the northern and southern hemispheres separately are given in Table2 (DS Using both procedures, interactive and automatic, five solar cy- method – rows 14 and 21; rLSQ method – rows 35 and 42). cles (20–24) from 1964 until 2016 were used to obtain the differ- Corresponding differential rotation profiles are shown in Fig.6. ential rotation parameters. Both solar hemispheres were treated More details are given in Sect. 6.3. together and separately. With only a CMD cutoff we obtained a sample of 41 125 (33 817 via DS and 7308 via rLSQ) calculated sidereal rotation 6. Discussion velocities. Rotational velocity outliers resulting from misidenti- fication of sunspot groups in subsequent images can be filtered 6.1. Automatic vs. interactive procedure, DS vs. rLSQ out by applying the velocity filter 8–19 deg/day (Brajša et al. method 2002b; Vršnak et al. 2003; Sudar et al. 2014, 2015). After se- If we compare the results for the year 2014 determined by the lecting only sidereal rotation velocities higher than 8 deg/day interactive and automatic procedures and by the DS method and lower than 19 deg/day, the number of calculated sidereal ro- (Table1, rows 1 and 2), for both di fferential rotation parameters tation velocities was reduced to 40 480 (33 213 via DS and 7267 the results are within the 1 common σ, which also holds for the via rLSQ). This means that 1.82% of the calculated velocities were eliminated using the DS method and 0.56% using the rLSQ 6 http://www.sidc.be/silso/news004

A72, page 4 of 10 I. Poljanciˇ c´ Beljan et al.: Solar differential rotation determined by the Kanzelhöhe data set

Table 2. Results of fittings of the KSO data to rotational velocity profile Eq. (2).

Row Method Cycle Time period Hemisphere ABNvl 1 int, DS 20 1964.8–1976.3 N+S 14.45 ± 0.02 −3.05 ± 0.27 5319 2 int, DS 21 1976.3–1986.7 N+S 14.50 ± 0.02 −2.47 ± 0.20 7486 3 int, DS 22 1986.7–1996.4 N+S 14.44 ± 0.02 −2.81 ± 0.17 8400 4 int, DS 23 1996.4–2008.9 N+S 14.50 ± 0.02 −2.54 ± 0.17 8204 5 aut, DS 24 2008.9–2016.3 N+S 14.48 ± 0.04 −2.61 ± 0.45 3701 6 int (all), DS 20–23 1964.8–2008.9 N+S 14.47 ± 0.01 −2.66 ± 0.10 29 409 7 int + aut (all), DS 20–24 1964.8–2016.3 N+S 14.47 ± 0.01 −2.66 ± 0.10 33 110 8 int, DS 20 1964.8–1976.3 N 14.45 ± 0.03 −3.17 ± 0.35 2845 9 int, DS 21 1976.3–1986.7 N 14.49 ± 0.03 −2.26 ± 0.29 3675 10 int, DS 22 1986.7–1996.4 N 14.43 ± 0.03 −3.13 ± 0.25 3882 11 int, DS 23 1996.4–2008.9 N 14.50 ± 0.03 −2.79 ± 0.25 3838 12 aut, DS 24 2008.9–2016.3 N 14.47 ± 0.06 −2.48 ± 0.68 1893 13 int (all), DS 20–23 1964.8–2008.9 N 14.47 ± 0.02 −2.83 ± 0.14 14 240 14 int + aut (all), DS 20–24 1964.8–2016.3 N 14.47 ± 0.02 −2.82 ± 0.14 16 133 15 int, DS 20 1964.8–1976.3 S 14.44 ± 0.03 −2.78 ± 0.45 2474 16 int, DS 21 1976.3–1986.7 S 14.52 ± 0.03 −2.67 ± 0.28 3811 17 int, DS 22 1986.7–1996.4 S 14.44 ± 0.03 −2.52 ± 0.24 4518 18 int, DS 23 1996.4–2008.9 S 14.49 ± 0.03 −2.33 ± 0.24 4366 19 aut, DS 24 2008.9–2016.3 S 14.49 ± 0.06 −2.74 ± 0.62 1808 20 int (all), DS 20–23 1964.8–2008.9 S 14.47 ± 0.02 −2.50 ± 0.14 15 169 21 int + aut (all), DS 20–24 1964.8–2016.3 S 14.47 ± 0.01 −2.51 ± 0.13 16 977 22 int, rLSQ 20 1964.8–1976.3 N+S 14.45 ± 0.03 −3.06 ± 0.33 1216 23 int, rLSQ 21 1976.3–1986.7 N+S 14.53 ± 0.03 −2.66 ± 0.26 1717 24 int, rLSQ 22 1986.7–1996.4 N+S 14.50 ± 0.02 −3.17 ± 0.20 1763 25 int, rLSQ 23 1996.4–2008.9 N+S 14.51 ± 0.03 −2.79 ± 0.22 1757 26 aut, rLSQ 24 2008.9–2016.3 N+S 14.52 ± 0.05 −2.60 ± 0.52 787 27 int (all), rLSQ 20–23 1964.8–2008.9 N+S 14.50 ± 0.01 −2.89 ± 0.12 6453 28 int + aut (all), rLSQ 20–24 1964.8–2016.3 N+S 14.50 ± 0.01 −2.87 ± 0.12 7240 29 int, rLSQ 20 1964.8–1976.3 N 14.46 ± 0.04 −3.26 ± 0.39 640 30 int, rLSQ 21 1976.3–1986.7 N 14.51 ± 0.04 −2.35 ± 0.38 836 31 int, rLSQ 22 1986.7–1996.4 N 14.50 ± 0.03 −3.50 ± 0.27 803 32 int, rLSQ 23 1996.4–2008.9 N 14.52 ± 0.04 −2.89 ± 0.34 813 33 aut, rLSQ 24 2008.9–2016.3 N 14.41 ± 0.07 −1.87 ± 0.87 405 34 int (all), rLSQ 20–23 1964.8–2008.9 N 14.50 ± 0.02 −3.01 ± 0.17 3092 35 int + aut (all), rLSQ 20–24 1964.8–2016.3 N 14.49 ± 0.02 −2.94 ± 0.17 3497 36 int, rLSQ 20 1964.8–1976.3 S 14.43 ± 0.05 −2.76 ± 0.57 576 37 int, rLSQ 21 1976.3–1986.7 S 14.54 ± 0.04 −2.94 ± 0.36 881 38 int, rLSQ 22 1986.7–1996.4 S 14.49 ± 0.03 −2.83 ± 0.29 960 39 int, rLSQ 23 1996.4–2008.9 S 14.50 ± 0.03 −2.71 ± 0.30 944 40 aut, rLSQ 24 2008.9–2016.3 S 14.66 ± 0.07 −3.59 ± 0.63 382 41 int (all), rLSQ 20–23 1964.8–2008.9 S 14.49 ± 0.02 −2.77 ± 0.17 3361 42 int + aut (all), rLSQ 20–24 1964.8–2016.3 S 14.51 ± 0.02 −2.81 ± 0.17 3743

Notes. Differential rotation parameters are calculated separately for solar cycles Nos. 20–24, for the part of the data set using the interactive procedure (covering solar cycles Nos. 20–23) – int (all), and for the whole data set using the interactive and automatic procedure together (covering solar cycles Nos. 20–24) – int+aut (all). All results are calculated separately via the DS and rLSQ methods covering the whole Sun (i.e., both hemispheres together) – N+S, the northern hemisphere – N, and the southern hemisphere – S. All calculations are exposed to ±58 deg CMD filter and 8–19 deg/day velocity filter. The sidereal parameters and their standard errors are expressed in deg/day. Nvl is the number of calculated rotation velocities. rLSQ results (Table1, rows 3 and 4). However, both procedures, It is obvious that the differential rotation profiles for the the interactive and the automatic, show large standard errors, es- rLSQ method coincide better (Fig.3, black and blue lines) be- pecially for parameter B. A comparison of the standard errors in cause the equatorial rotation velocities are almost the same and Table1 (rows 1 and 2, rows 3 and 4) between the two procedures the gradients of the solar rotation do not differ as in the DS case. reveals that the automatic procedure is less accurate but can pro- In the case of the automatic procedure, the rLSQ method for cal- cess more data. The application of the interactive procedure is culating rotational velocities (Fig.3, blue line) is much more re- rather time consuming and the results presented here are a com- liable than the DS method (Fig.3, red line), which can be clearly bination of the efforts of two co-authors over almost two years seen by a comparison of the corresponding differential rotation of measurements. parameters in Table1 (rows 2 and 4). For example, parameter B

A72, page 5 of 10 A&A 606, A72 (2017)

Table 3. Values of differential rotation parameters A and B and their standard errors (in deg/day) collected from different sources using only sunspots and sunspot groups as tracers.

Row Data set Time Hemisphere A B References 1 GPRa 1874–1976 N+S 14.551 ± 0.006 −2.87 ± 0.06 1 2 GPRa 1874–1902 N+S 14.63 ± 0.01 −2.70 ± 0.16 2 3 GPRa 1879–1975 N+S 14.522 ± 0.005 −2.66 ± 0.04 3 4 GPRb 1880–1976 N+S 14.37 ± 0.01 −2.59 ± 0.16 4 5 GPRa 1883–1893 N+S 14.63 ± 0.0− −2.69 ± 0.0− 5 6 GPRa 1940–1968 N+S 14.53 ± 0.01 −2.83 ± 0.08 6 7 GPRa 1948–1976 N+S 14.52 ± 0.0− −2.84 ± 0.0− 5 8 EGRa 1878–2011 N+S 14.49 ± 0.01 −2.64 ± 0.05 7 9 EGRa 1976–2002 N+S 14.457 ± 0.009 −2.17 ± 0.07 3 10 EGRc 1874–1996 N+S 14.531 ± 0.003 −2.747 ± 0.048 8 11 CSc 1853–1893 N+S 14.475 ± 0.011 −2.710 ± 0.165 8 12 Spörera 1883–1893 N+S 14.50 ± 0.0− −2.41 ± 0.0− 5 13 Abastumanid 1950–1990 N+S 14.73 ± 0.06 −2.07 ± 0.51 9 14 Mt. Wilsond 1921–1982 N+S 14.522 ± 0.004 −2.84 ± 0.04 10 15 Mt. Wilsona 1921–1982 N+S 14.393 ± 0.010 −2.95 ± 0.09 10 16 DPDa 1974–2016 N+S 14.50 ±0.01 −2.54 ± 0.07 11 17 KSOc 1970–1979 N+S 14.27 ± 0.02 −1.84 ± 0.12 12 18 KSOc 1947–1981 N+S 14.38 ± 0.01 −2.57 ± 0.07 13 19 KSOa 1948–1976 N+S 14.35 ± 0.0− −2.73 ± 0.0− 5 20 KSOa 1964–2016 N+S 14.47 ± 0.01 −2.66 ± 0.10 15 21 KSOa 1964–2016 N+S 14.50 ± 0.01 −2.87 ± 0.12 16 22 GPRa 1874–1976 N 14.54 ± 0.01 −2.88 ± 0.08 1 23 GPRa 1874–1902 N 14.65 ± 0.02 −2.89 ± 0.16 2 24 GPRa 1879–1975 N 14.531 ± 0.005 −2.63 ± 0.06 3 25 GPRa 1940–1968 N 14.51 ± 0.01 −2.69 ± 0.11 6 26 EGRa 1976–2002 N 14.462 ± 0.015 −2.13 ± 0.10 3 27 EGRc 1874–1996 N 14.53 ± 0.01 −2.69 ± 0.07 8 28 CSc 1853–1893 N 14.48 ± 0.02 −3.02 ± 0.26 8 29 KSOc 1947–1981 N 14.38 ± 0.01 −2.70 ± 0.09 13 30 KSOc 1964–1976 N 14.40 ± 0.02 −2.75 ± 0.24 14 31 KSOa 1964–2016 N 14.47 ± 0.02 −2.82 ± 0.14 17 32 KSOa 1964–2016 N 14.49 ± 0.02 −2.94 ± 0.17 18 33 GPRa 1874–1976 S 14.56 ± 0.01 -2.85 ± 0.09 1 34 GPRa 1874–1902 S 14.61 ± 0.02 −2.56 ± 0.16 2 35 GPRa 1879–1975 S 14.517 ± 0.005 −2.68 ± 0.05 3 36 GPRa 1940–1968 S 14.55 ± 0.01 −3.00 ± 0.13 6 37 EGRa 1976–2002 S 14.448 ± 0.015 −2.20 ± 0.10 3 38 EGRc 1874–1996 S 14.54 ± 0.01 −2.81 ± 0.07 8 39 CSc 1853–1893 S 14.48 ± 0.02 −2.51 ± 0.22 8 40 KSOc 1947–1981 S 14.38 ± 0.01 −2.34 ± 0.11 13 41 KSOc 1964–1976 S 14.37 ± 0.02 −2.48 ± 0.27 14 42 KSOa 1964–2016 S 14.47 ± 0.01 −2.51 ± 0.13 19 43 KSOa 1964–2016 S 14.51 ± 0.02 −2.81 ± 0.17 20

Notes. N denotes the northern hemisphere, S denotes the southern hemisphere, and N+S both hemispheres together. In Col. 2: GPR – Greenwich photoheliographic results, EGR – extended Greenwich results, KSO – Kanzelhöhe Observatory for Solar and Environmental Research data, CS – Carrington & Spörer data, DPD – Debrecen photoheliographic results. Type of tracers: (a) sunspot groups; (b) stable recurrent sunspot groups; (c) sunspots and sunspot groups; (d) sunspots. References. (1) Balthasar et al.(1986b); (2) Arévalo et al.(1982); (3) Javaraiah(2003); (4) Brajša et al.(2002a); (5) Balthasar & Fangmeier (1988); (6) Balthasar & Wöhl(1980); (7) Sudar et al.(2014); (8) Pulkkinen & Tuominen(1998); (9) Khutsishvili et al.(2002); (10) Howard et al. (1984); (11) Sudar et al.(2017); (12) Lustig(1982); (13) Lustig(1983); (14) Lustig(1983), cycle 20; (15) present work, Table2, row 7, DS; (16) present work, Table2, row 28, rLSQ; (17) present work, Table2, row 14, DS; (18) present work, Table2, row 35, rLSQ; (19) present work, Table2, row 21, DS; (20) present work, Table2, row 42, rLSQ.

A72, page 6 of 10 I. Poljanciˇ c´ Beljan et al.: Solar differential rotation determined by the Kanzelhöhe data set

0 10 20 30 40 50 rotation velocities for the whole time period (daily shift – DS and robust linear least-squares fit – rLSQ) can be considered al- 14.4 14.4 most identical. 14.2 14.2

14.0 14.0 6.2. Comparison with other data sets )

da y 13.8 13.8 In Poljanciˇ c´ et al.(2011) a comparison of sunspot position mea-

deg / 13.6 13.6 surements for several data sets was made, separately for single

KSO, 1970 - 1979, Lustig (1982), / ( H and J and complex sunspot groups types. The mean absolute ω 13.4 Table 3, row 17 13.4 KSO, 1947 - 1981, Lustig (1983), differences between the KSO and DPD positions are lower in all T ab l e 3 , row 18 13.2 KSO, 1948 - 1976, Balthasar & Fangmeier (1988), 13.2 cases – except for the longitude difference for complex sunspot Table 3, row 19 groups – than the mean absolute difference between the KSO 13.0 KSO, 1964 - 2016, present work, DS, 13.0 T ab l e 3 , row 20 and GPR positions. As the KSO and GPR comparison was made KSO, 1964 - 2016, present work, rLSQ, 12.8 Table 3, row 21 12.8 for the year 1972, it may be possible that the KSO data is influ- enced by the erroneous solar image radii (Balthasar et al. 1984), 0 10 20 30 40 50 thus causing a larger difference than that between the KSO and b / deg the DPD. The KSO and DPD comparison was made for the year Fig. 4. Differential rotation profiles calculated by different authors for 1993, and was no longer influenced by the previously mentioned the KSO data set and both hemispheres together (see corresponding unwanted effect, because the KSO observing and reduction pro- rows in Table3). Sidereal rotation velocity is denoted by ω and heli- cedures were improved after the 1980s (Balthasar et al. 1984). ographic latitude by b. We compared the differential rotation parameters derived in the present work from KSO for both the DS and the rLSQ methods (Table3, rows 20 and 21) with values for the EGR calculated via the DS method yields B = (−0.98 ± 1.11) deg/day, (Sudar et al. 2014; Table3, row 8) and values for the DPD while for the rLSQ method B = (−3.40 ± 1.32) deg/day. If rel- (Sudar et al. 2017; Table3, row 16), which were also obtained ative standard errors are compared, the rLSQ method yields a by tracing sunspot groups and covering almost the same time pe- value that is three times smaller than the corresponding value for riod. Values for the differential rotation parameter A are within the DS method. the 1 common σ, while values for the differential rotation pa- However, if we look at the results obtained by applying only rameter B match within the 2 common σ. It is well known that the interactive procedure (Table2, rows 6, 13, 20, 27, 34, 41) and the GPR represents a homogeneous data set with high accu- compare them with the results of the interactive and automatic racy (Balthasar & Wöhl 1980; Arévalo et al. 1982; Willis et al. procedures together covering the whole time period (Table2, 2013a,b; Erwin et al. 2013), as well as the DPD as its continua- rows 7, 14, 21, 28, 35, 42), we see that the inclusion of the results tion (Sudar et al. 2017). Therefore, we feel free to conclude that calculated by the automatic procedure, for solar cycle No. 24, the KSO data used in this paper (from 1964 till present) is of does not affect the “whole time period” result. In other words, the comparable accuracy to the DPD or GPR data sets, and suitable int (all) and int + aut (all) results from the corresponding rows in for the investigation of the solar rotation, although there is some Table2, which are listed above, are almost identical in all cases indication of lower accuracy in the past, especially before the for both the DS and rLSQ methods and for both hemispheres. 1960s (Balthasar et al. 1984). Thus, when the rotation velocities calculated by the automatic Except for the already mentioned overlapping with procedure are combined with the rotation velocities calculated Sudar et al.(2014, 2017) rotation results (Table3, rows 8 and by the interactive procedure for longer periods of time, the effect 16), in most of the other studies listed in Table3 the di fferential of the lower accuracy of the automatic procedure has no statis- rotation parameters are in agreement with our results Table3, tically significant influence on the final best fit differential rota- rows 20 and 21); the exceptions are the GPR result for rec- tion profile. Hence, we combine both procedures if we analyse curent sunspot groups (Brajša et al. 2002a; Table3, row 4); the the whole period from 1964 to 2016. No separation between an Mt. Wilson result concerning only sunspot groups as tracers interactive procedure before 2009 and an automatic procedure (Howard et al. 1984; Table3, row 15); the Abastumani result after 2009 has to be made. (Khutsishvili et al. 2002; Table3, row 13); and the KSO results For white light images recorded on photographic film mate- (Lustig 1982, 1983; Balthasar & Fangmeier 1988; Table3, rows rial before 2007, which are scanned, the automatic procedure did 17–19). not yield reliable results, and we were forced to process this data According to Ruždjak et al.(2004, 2005) recurrent sunspots using the interactive procedure and sunspot drawings. So, for and recurrent sunspot groups show a slower rotation, which may the years before 2007, it is possible to apply the automatic pro- be the reason for the low rotation values of the GPR given by cedure after elimination of the possible errors that appear during Brajša et al.(2002a; Table3, row 4). It is interesting that the the scanning procedure. However, for the years after 2007, the lower value of the equatorial velocity A was also obtained for the automatic procedure represents a useful tool for processing the Mt. Wilson data for sunspot groups (Howard et al. 1984; Table3, white light images. row 15), but it is not clear what could be the cause. A higher If we look at Table2 and compare the DS and rLSQ measure- value of the equatorial velocity, as well as the lower value of the ments derived for the whole periods (both hemispheres – rows 7 differential rotation gradient was obtained with the Abastumani and 28; northern hemisphere – rows 14 and 35; southern hemi- data (Khutsishvili et al. 2002; Table3, row 13) compared to our sphere – rows 21 and 42), the errors for both differential rota- analysis (Table3, rows 20 and 21). This result, however, shows tion parameters in all cases (northern, southern, or both hemi- higher standard errors in comparison to other results in Table3. spheres) are within the 1 common σ. In some cases there are The equatorial rotation rates A derived from the KSO data very small (i.e., almost negligible) differences. This means that (Lustig 1982, 1983; Balthasar & Fangmeier 1988; Table3, rows the results of the two different methods used to determine the 17–19) for the years before the 1980s are significantly lower

A72, page 7 of 10 A&A 606, A72 (2017)

0 5 10 15 20 25 30

14.5 14.5

14.4 14.4

14.3 14.3

14.2 14.2 ) da y /

de g 14.1 14.1

/ ( ω

14.0 GPR, 1874 - 1976, Balthasar et al. (1986), Table 3, row 1 14.0 GPR, 1879 - 1975, Javaraiah (2003), Table 3, row 3

GPR, 1940 - 1968, Balthasar & Wöhl (1980), Table 3, row 6 EGR, 1878 - 2011, Sudar et al. (2014), Table 3, row 8 13.9 13.9 EGR, 1976 - 2002, Javaraiah (2003), Table 3, row 9 EGR, 1874 - 1996, Pulkkinen & Tuominen (1998), Table 3, row 10 M t. W il s on , 1921 - 1982 , H o w a rd e t a l. (1984), Table 3 , r o w 1 5 13.8 DPD, 1974 - 2016, Sudar et al. (2017), Table 3, row 16 13.8 KSO, 1947 - 1981, Lustig (1983), Table 3, row 18 KSO, 1964 - 2016, present work, DS, Table 3, row 20 KSO, 1964 - 2016, present work, rLSQ, Table 3, row 21 13.7 13.7

0 5 10 15 20 25 30 b / deg Fig. 5. Differential rotation profiles calculated by different authors for several data sets (GPR, EGR, Mt. Wilson, DPD, KSO) and both hemispheres together (see corresponding rows in Table3). Sidereal rotation velocity is denoted by ω and heliographic latitude by b.

than our analysis (Table3 – rows 20 and 21) by about 0.15 deg. the KSO data set has been improved, not only for the period after A large difference between differential rotation parameters B is the 1980s, but indeed from 1960 onwards. also noticed. Balthasar & Fangmeier(1988) analysed the GPR data for the same time period (Table3, row 7) and also noted the disagreement with the KSO data, confirming that the KSO data 6.3. North–south asymmetry of the solar rotation before the 1980s are affected by some systematic errors, already examined in Balthasar et al.(1984). If we compare the DS measurements derived for the whole time Since our investigations were conceived as a continuation of period (1964–2016) and for the northern and the southern hemi- the investigations done in Lustig(1983), it is important to com- spheres separately (Table2, rows 14 and 21), we see that the pare them also for the overlapping period of time (1964–1976, measurements for differential rotation parameter A are within the solar cycle No. 20). Lustig(1983) used the Stonyhurst disk over- 1 common σ, i.e., the difference is very small and statistically laying technique to determine the heliographic coordinates of insignificant. Differential rotation parameter B shows the differ- sunspot groups by superimposing the disk on sunspot drawing. ence between the two solar hemispheres, statistically significant For solar cycle No. 20, Lustig(1983) provides the solar rota- on 2 common σ level (Table2, rows 14 and 21). Concerning tion parameters only for the northern and southern hemispheres. the rLSQ measurements (Table2, rows 35 and 42), we see that They are listed in Table3 (rows 30 and 41). The corresponding the measurements for both differential rotation parameters are solar rotation parameters derived in the present paper are listed within the 1 common σ. in Table2 (rows 8 and 15 for DS, rows 29 and 36 for rLSQ). All In Lustig(1983), statistically insignificant di fferences were these rotation profiles are shown in Fig.7. The results for both discovered for equatorial velocities (the parameter A) between differential rotation parameters and both methods, the DS and the northern and southern hemispheres, while the difference in the rLSQ, are within the 1 common σ. Only the differential ro- the gradients of the differential rotation (the parameter B) is tation parameter A for the DS method and southern hemisphere 0.36 deg/day. It is interesting that in present work for the DS falls within the 2 common σ with the corresponding result from method this difference remains almost the same (BN − BS = Lustig(1983). This comparison is evidence that the reliability of 0.31 deg/day).

A72, page 8 of 10 I. Poljanciˇ c´ Beljan et al.: Solar differential rotation determined by the Kanzelhöhe data set

b / deg b / deg 0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50 14.6 14.6 14.4 14.4 rLSQ (N), Table 2, row 35 14.2 Northern 14.2 rLSQ (S), Table 2, row 42 hemisphere 14.0 14.0 14.6

) ) 14.4 13.8 14.6 13.8 14.2 13.6 13.6 14.4 13.4 14.0 deg / da y deg / da y

13.4 ) 13.2 / ( 14.2 / ( 13.8 ω ) ω Southern 13.2 13.0 14.0 13.6 hemisphe re 13.0 13.4 12.8 13.8 de g/ da y

deg / da y 12.6 DS (N), Table 2, row 14 12.8 / ( 13.2 13.6 ω / ( DS (S), Table 2, row 21 12.4

ω 13.0 13.4 12.8 present work, int, DS, cycle 20 (Table 2, row 8 and 15) 13.2 present work, int, rLSQ, cycle 20 (Table 2, row 29 and 36) 12.6 Lustig 1983, cycle 20 (Table 3, row 30 and 41) 13.0 12.4 12.8 0 5 10 15 20 25 30 35 40 45 50 b / deg 0 5 10 15 20 25 30 35 40 45 50 b / deg Fig. 7. Comparison between the present paper and Lustig(1983) dif- Fig. 6. North – south asymmetry for the whole time period 1964–2016 ferential rotation profiles calculated for solar cycle No. 20 (1964–1976) and for DS and rLSQ methods (see also the corresponding rows in and for the northern and southern hemisphere separately, see corre- Table2). Sidereal rotation velocity is denoted by ω and heliographic sponding rows in Tables2 and3. The sidereal rotation velocity is de- latitude by b. N – northern hemisphere, S – southern hemisphere. noted by ω and heliographic latitude by b.

For both methods, the DS and the rLSQ, the southern solar the automatic procedure are combined with the rotation veloc- hemisphere rotates a little bit faster, which can also be clearly ities calculated by the interactive procedure for longer periods seen in Fig.6. Therefore, a barely noticeable north–south asym- of time, the effect of the lower accuracy of the automatic pro- metry is observed for the whole time period 1964–2016 in the cedure has no statistically significant influence on the final best present work. fit differential rotation profile. This allows us to discuss interac- tive procedure before 2009 and automatic procedure after 2009 7. Conclusions together as if only one procedure has been applied. The two different methods used for the determination of We have determined the solar differential rotation by tracing the rotation velocities (daily shift – DS and robust linear least- sunspot groups during the period 1964–2016. We have used two squares fit – rLSQ) were also compared for the whole time pe- procedures to determine the heliographic positions: an interac- riod, and showed 1 common σ coincidence for both differential tive one on the KSO sunspot drawings (1964–2008, solar cy- rotation parameters. cles Nos. 20–23) and an automatic one on the KSO white light Inspection for the north – south asymmetry showed that for images (2009–2016, solar cycle No. 24). For the first time, the both methods, the DS and the rLSQ, the southern solar hemi- whole solar cycle No. 21 was investigated with the KSO data, sphere rotates a little bit faster. Thereby, a barely noticeable as well as solar cycles Nos. 22–24. Applying a CMD cutoff at north – south asymmetry is observed for the whole time period ± 58 deg we obtained a sample of 12 152 sunspot groups which 1964–2016 in the present paper, very similar to previous obser- correspond to approximately 70 000 individual sunspot posi- vations for the time period 1947–1981 (Lustig 1983). tions. The synodic angular rotation velocities were determined using two different methods, the DS and rLSQ, and then con- The comparison of calculated rotation profiles derived from verted to the sidereal velocities. A sample of 41 125 calculated the KSO data in this work and from all other studies collected sidereal rotation velocities was obtained. Calculated sidereal ve- from different sources yielded a conclusion about the KSO ac- locity values have been used in the least-squares fitting to the curacy after the 1960s. The KSO data used in this paper for the solar differential rotation law. time period from 1964 to the present is of a comparable accu- The interactive and automatic procedures comparison was racy with the DPD or GPR data sets, and suitable for the inves- performed for the year 2014. It yielded 1 common σ coinci- tigation of the solar rotation. We cannot guarantee that for the dence for both differential rotation parameters and both the DS years prior to 1960 the results would be of the same accuracy and rLSQ methods. However, in both procedures, interactive and because there are some indications about lower accuracy in the automatic, there are large standard errors for the differential ro- past (Balthasar et al. 1984). tation parameters, especially for the differential rotation param- The quality of the KSO sunspot drawings has gradually in- eter B. If the comparison were performed for the entire cycle, creased over the last 50 yr. In general, the interactive procedure it would yield better results as the number of calculated veloc- of position determination is fairly accurate, but has the draw- ities and latitudinal coverage is higher in that case. For the au- back of being very time consuming, so the much faster automatic tomatic procedure the standard errors are even larger, indicat- procedure of position determination was developed. However, in ing lower accuracy of the automatic procedure in comparison to the case of the automatic procedure, the rLSQ method for cal- the interactive one. Also, in the case of automatic procedure, the culating rotational velocities is much more reliable than the DS rLSQ method for calculating rotational velocities is much more method. reliable than the DS method. For the test data from 2014, the The main conclusion of this paper is that KSO provides a rLSQ method yields a relative standard error for the differen- valuable data set with satisfactory accuracy. Although there were tial rotation parameter B that is three times smaller than that of some systematic errors in the past, this data set is well suited to the DS method. But, when the rotation velocities calculated by long-term studies and therefore still continued.

A72, page 9 of 10 A&A 606, A72 (2017)

We plan to continue our analysis of the KSO data. The next Javaraiah, J., & Ulrich, R. K. 2006, Sol. Phys., 237, 245 step is presenting the temporal variation of the differential ro- Javaraiah, J., Bertello, L., & Ulrich, R. K. 2005, Sol. Phys., 232, 25 tation and the relationship between the solar rotation and activ- Khutsishvili, E. V., Gigolashvili, M. S., & Kvernadze, T. M. 2002, Sol. Phys., 206, 219 ity, as well as an analysis of meridional motions and horizontal Lustig, G. 1982, A&A, 106, 151 Reynolds stress. Lustig, G. 1983, A&A, 125, 355 Lustig, G., & Dvorak, R. 1984, A&A, 141, 105 Acknowledgements. This work was supported in part by the Croatian Science Lustig, G., & Hanslmeier, A. 1987, A&A, 172, 332 Foundation under the project 6212 “Solar and Stellar Variability” and in part Lustig, G., & Wöhl, H. 1991, A&A, 249, 528 by the University of Rijeka under project number 13.12.1.3.03. It has also re- Meeus, J. 1991, Astronomical Algorithms (Willmann-Bell) ceived funding from the SOLARNET project (312495, 2013–2017), which is an Olemskoy, S. V., & Kitchatinov, L. L. 2005, Astron. Lett., 31, 706 Integrated Infrastructure Initiative (I3) supported by FP7 Capacities Programme. Otruba, W. 2006, Sun and Geosphere, 1, 21 I.P.B. thanks the Kanzelhöhe Solar Observatory for the hospitality during her Poljanciˇ c,´ I., Brajša, R., Ruždjak, D., et al. 2010, Sun and Geosphere, 5, 52 stay at Kanzelhöhe. Poljanciˇ c,´ I., Brajša, R., Hržina, D., et al. 2011, Central European Astrophysical Bulletin, 35, 59 Poljanciˇ c´ Beljan, I., Jurdana-Šepic,´ R., Cargonja,ˇ M., et al. 2014, Central European Astrophysical Bulletin, 38, 87 References Pötzi, W. 2010, Central European Astrophysical Bulletin, 34, 1 Pötzi, W., Veronig, A. M., Temmer, M., et al. 2016, Sol. Phys., 291, 3103 Arévalo, M. J., Gomez, R., Vázquez, M., Balthasar, H., & Wöhl, H. 1982, A&A, Pulkkinen, P., & Tuominen, I. 1998, A&A, 332, 755 111, 266 Roša, D., Brajša, R., Vršnak, B., & Wöhl, H. 1995, Sol. Phys., 159, 393 Balthasar, H., & Fangmeier, E. 1988, A&A, 203, 381 Ruždjak, D., Ruždjak, V., Brajša, R., & Wöhl, H. 2004, Sol. Phys., 221, 225 Balthasar, H., & Wöhl, H. 1980, A&A, 92, 111 Ruždjak, D., Brajša, R., Sudar, D., & Wöhl, H. 2005, Sol. Phys., 229, 35 Balthasar, H., Lustig, G., & Wöhl, H. 1984, Sol. Phys., 91, 55 SILSO World Data Center 2015, 2014: maximum year for , Balthasar, H., Lustig, G., Wöhl, H., & Stark, D. 1986a, A&A, 160, 277 Royal Observatory of Belgium, Brussels, http://www.sidc.be/silso/ Balthasar, H., Vázquez, M., & Wöhl, H. 1986b, A&A, 155, 87 news004, online accessed 7 April 2015 Baranyi, T., Gyori,˝ L., & Ludmány, A. 2016, Sol. Phys., 291, 3081 Skokic,´ I., Brajša, R., Roša, D., Hržina, D., & Wöhl, H. 2014, Sol. Phys., 289, Brajša, R., Wöhl, H., Vršnak, B., et al. 2002a, Sol. Phys., 206, 229 1471 Brajša, R., Wöhl, H., Vršnak, B., et al. 2002b, A&A, 392, 329 Stix, M. 2004, The sun: an introduction (Berlin: Springer-Verlag) Brajša, R., Ruždjak, D., & Wöhl, H. 2006, Sol. Phys., 237, 365 Sudar, D., Skokic,´ I., Ruždjak, D., Brajša, R., & Wöhl, H. 2014, MNRAS, 439, Brajša, R., Wöhl, H., Ruždjak, D., et al. 2007, Astron. Nachr., 328, 1013 2377 Brajša, R., Wöhl, H., Hanslmeier, A., et al. 2009, A&A, 496, 855 Sudar, D., Skokic,´ I., Brajša, R., & Saar, S. H. 2015, A&A, 575, A63 Erwin, E. H., Coffey, H. E., Denig, W. F., et al. 2013, Sol. Phys., 288, 157 Sudar, D., Brajša, R., Skokic,´ I., Poljanciˇ c´ Beljan, I., & Wöhl, H. 2017, Gilman, P., & Howard, R. 1984, ApJ, 283, 385 Sol. Phys., 292, 86 Gupta, S. S., Sivaraman, K. R., & Howard, R. F. 1999, Sol. Phys., 188, 225 Temmer, M., Rybák, J., Bendík, P., et al. 2006, A&A, 447, 735 Gyori,˝ L., Ludmány, A., & Baranyi, T. 2017, MNRAS, 465, 1259 Veronig, A. M., & Pötzi, W. 2016, in Coimbra Solar Physics Meeting: Hanslmeier, A., & Lustig, G. 1986, A&A, 154, 227 Ground-based Solar Observations in the Space Instrumentation Era, eds. Hathaway, D. H., & Wilson, R. M. 1990, ApJ, 357, 271 I. Dorotovic, C. E. Fischer, & M. Temmer, ASP Conf. Ser., 504, 247 Howard, R. 1984, ARA&A, 22, 131 Vršnak, B., Brajša, R., Wöhl, H., et al. 2003, A&A, 404, 1117 Howard, R., Gilman, P. I., & Gilman, P. A. 1984, ApJ, 283, 373 Watson, F., & Fletcher, L. 2011, in Physics of Sun and Star Spots, eds. D. Prasad Hržina, D., Roša, D., Hanslmeier, A., Ruždjak, V., & Brajša, R. 2007, Central Choudhary, & K. G. Strassmeier, IAU Symp., 273, 51 European Astrophysical Bulletin, 31 Willis, D. M., Coffey, H. E., Henwood, R., et al. 2013a, Sol. Phys., 288, 117 Huber, P. J. 1981, Robust Statistics (New York: John Wiley & Sons, Inc.) Willis, D. M., Henwood, R., Wild, M. N., et al. 2013b, Sol. Phys., 288, 141 Javaraiah, J. 2003, Sol. Phys., 212, 23 Zuccarello, F., & Zappalá, R. 2003, Astron. Nachr., 324, 464

A72, page 10 of 10 4.2 Paper II: A relationship between the solar rotation and activity in the period 1998–2006 analysed by tracing small bright coronal structures in SOHO-EIT images

The following article is published in Astronomy and Astrophysics, Volume 534, Issue A17, pp. 1-6 (2011). It is reproduced by the permission of ESO. The .pdf document is available in the online version of this journal and hereafter. My own contribution to this work was 20%.

The main aims of this paper are presented in Sec. 1. The main conclusions of this paper are listed here. The first among them is the conclusion about the dependence of the solar rotation on the phase of the solar cycle. This dependence is obvious for the solar rotation parameter A, whilst the results are not conclusive for parameter B. During the maximum of the and just after it, the equatorial solar rotation velocity A was lower than in other phases of the cycle, when there was less activity. This is consistent with other observational findings, obtained by different tracers and methods. Second conclusion is connected with the investigation of the relationship between the solar rotation and activity. Solar rotation is presented by the differential rotation parameters A and B, while solar activity is expressed by the monthly relative sunspot number (monthly SN), the smoothed monthly relative sunspot number (monthly SSN), the yearly relative sunspot number (yearly SN), and the interdiurnal variability (IDV) index. The statistical analysis for the annual values of the solar sidereal rotation param- eters A and B have been determined as a function of the solar activity (SN and SSN) and also as a function of the interplanetary magnetic field strength, derived from the IDV index. The statistically significant correlation was found for the solar rotation pa- rameter A, whilst a very low and insignificant correlation was obtained for the rotation parameter B.

48 A&A 534, A17 (2011) Astronomy DOI: 10.1051/0004-6361/201014357 & c ESO 2011 Astrophysics

A relationship between the solar rotation and activity in the period 1998–2006 analysed by tracing small bright coronal structures in SOHO-EIT images

R. Jurdana-Šepic´1 ,R.Brajša2,H.Wöhl3,A.Hanslmeier4, I. Poljanciˇ c´1,L.Svalgaard5, and S. F. Gissot6

1 Physics Department, University of Rijeka, Omladinska 14, 51000 Rijeka, Croatia e-mail: [jurdana,ipoljancic]@phy.uniri.hr 2 Hvar Observatory, Faculty of Geodesy, University of Zagreb, Kaciˇ ceva´ 26, 10000 Zagreb, Croatia e-mail: [email protected] 3 Kiepenheuer-Institut für Sonnenphysik, Schöneckstr. 6, 79104 Freiburg, Germany e-mail: [email protected] 4 Institut für Physik, IGAM, Universität Graz, Universitätsplatz 5, 8010 Graz, Austria e-mail: [email protected] 5 Stanford University, HEPL, Stanford, CA 94305-4085, USA e-mail: [email protected] 6 SIDC, Royal Observatory of Belgium, Ave. Circulaire 3, 1180 Brussels, Belgium e-mail: [email protected]

Received 5 March 2010 / Accepted 4 August 2011

ABSTRACT

Aims. The study aims to find a relationship between the rotation of the small bright coronal structures (SBCS) described by the solar rotation parameters and indices of solar activity on monthly and yearly temporal scales. Methods. We analyse precise measurements of the solar differential rotation determined by tracing SBCS in SOHO-EIT images and compare the derived solar rotation parameters with the status of solar activity in the period 1998−2006. Full-disc solar images obtained with the Extreme ultraviolet Imaging Telescope (EIT) on board the Solar and Heliospheric Observatory (SOHO) were used to analyse solar differential rotation determined by tracing SBCS. An automatic method to identify and track the SBCS in EIT full-disc images with a six hour cadence is applied. We performed a statistical analysis of the monthly and yearly values of solar sidereal rotation velocity parameters A and B (corresponding to the equatorial rotation velocity and the gradient of the solar differential rotation, respectively) as a function of various solar activity indices. Results. The dependence of the solar rotation on the phase of the solar cycle was found. It is clearly visible for the solar rotation parameter A, whilst the results are not conclusive for parameter B. The relationship between the solar rotation and activity, expressed by the monthly relative sunspot number, the smoothed monthly relative sunspot number, the yearly relative sunspot number, and the interdiurnal variability (IDV) index was investigated. The statistically significant correlation was found for the solar rotation parameter A, whilst a very low and insignificant correlation was obtained for the rotation parameter B. Conclusions. During the maximum of the solar cycle 23 and just after it, the equatorial solar rotation velocity was lower than in other phases of the cycle, when there was less activity. This is consistent with other observational findings, obtained by different tracers and methods. Key words. Sun: rotation – Sun: corona – Sun: activity

1. Introduction sets available from various observatories. For the investigation of possible variations in the solar rotation, the Greenwich data Measurements of the solar differential rotation provide precisely set was used by Balthasar & Wöhl (1980), Arévalo et al. (1982), determined observational constraint on theoretical models of the and Balthasar et al. (1986); the Kanzelhöhe data set by Lustig solar convection zone and the solar MHD dynamo. The conser- (1983); the Mt. Wilson data set by Gilman & Howard (1984)and vation of angular momentum, its transport to maintain the dif- by Hathaway & Wilson (1990); the Mitaka data set by Kambry ferential rotation profile, and an interplay with magnetic field & Nishikawa (1990); the Suzuka data set by Suzuki (1998); the are very important for understanding physical properties of the extended Greenwich data set by Pulkkinen & Tuominen (1998), Sun and solar-type stars. Many indications of cycle-related vari- Javaraiah (2003), Zuccarello & Zappalá (2003), Javaraiah et al. ations in the solar rotation have been reported using differ- (2005), Javaraiah & Ulrich (2006), and Brajša et al. (2006, ent tracers and methods (e.g., Howard 1984; Schröter 1985; 2007); the Kodaikanal data set by Gupta et al. (1999); and the Snodgrass 1992; Stix 2002). Most often, sunspots and sunspot Abastumani data set by Khutsishvili et al. (2002). Also, tem- groups were used as tracers mainly because of the long term data poral variations in the solar rotation were studied using other Article published by EDP Sciences A17, page 1 of 6 A&A 534, A17 (2011) features, e.g., magnetic fields (Sheeley et al. 1987; Brajša et al. monthly smoothed and yearly values) and the strength of the in- 1992; Sheeley et al. 1992), coronal holes (Nash et al. 1988), ob- terplanetary magnetic field represented by the interdiurnal vari- servations of the corona in the green line (Rybák 1994; Altrock ability (IDV) index (Svalgaard & Cliver 2005). 2003), Hα filaments (Japaridze & Gigolashvili 1992; Brajša et al. 1997; Gigolashvili et al. 2003), microwave low-brightness- temperature regions (Brajša et al. 1997, 1999, 2000), and faculae 2. The SOHO-EIT data set and the automatic (Meunier et al. 1997). Changes in the solar rotation were also method of data reduction investigated by the helioseismological method using the GONG and SOHO-MDI data at the base of the convective zone (Howe Several thousand full-disc full resolution solar filtergrams et al. 2000a,b, 2001) and at high latitudes (Antia & Basu 2001). recorded in the Fe XV line at the wavelength of 28.4 nm Recently, the time-series analysis was applied to investigate tem- taken with the Extreme Ultraviolet Imaging Telescope (EIT) poral changes of the solar rotation, especially using the radio in the years 1998−2006 were analysed. In these images SBCS data (Heristchi & Mouradian 2009; Chandra et al. 2009). Various were identified and used as tracers for the solar rotation de- frequencies correspond to various altitudes in the solar corona termination. The detailed description of the data set and the and different methods have been used (e.g., the Sun analysed as automatic method of data reduction are presented in Paper I a star with no latitudinal dependence, autocorrelation analysis (and references therein). The total amount of reduced data ex- with a time lag, etc.). However, it seems that up to now either a ceeds 50 000 triplets of SBCS (Paper I), which is of the same very weak or no correlation between the coronal rotational pe- order of magnitude as the number of pairs of daily position riod and the sunspot activity can be verified (Vats et al. 2010; measurements of sunspot groups analysed in the Greenwich Chandra & Vats 2011). This result refers to the height of about Photoheliographic Results by Balthasar et al. (1986). In the 60 000 km and the observing frequency of 2.8 GHz. present work we use the monthly and yearly parameters of the ff If we adopt the working hypothesis of the corrrelation be- solar di erential rotation published in Paper I and compare them tween the solar rotation and activity, then the strongest ef- with various solar activity proxies. fect should have been observed during the Maunder minimum (1645−1715). In spite of many investigations (e.g., Eddy et al. 1976; Abarbanell & Wöhl 1981; Ribes & Nesme-Ribes 1993; 3. Results Mendoza 1999; Vaquero et al. 2002; Casas et al. 2006), the re- 3.1. Temporal evolution sults are still not conclusive, owing the general uncertainty of measurements in those times and the different time subinter- As usually, the solar differential rotation is represented by vals when particular observations were performed (just before 2 4 or at the very beginning of the Maunder minimum, during the ω (ψ) = A + B sin ψ + C sin ψ, (1) Maunder minimum, during the deep Maunder minimum, and just after the regular cycles started again). In addition, new dis- where ω represents the angular rotation velocity expressed coveries and reductions of the observations from that time have in deg/day, ψ the heliographic latitude, and A−C the solar ro- often been reported. However, there is a strong evidence that tation parameters. In the present work we put C = 0tocompare sunspots rotated faster at low latitudes with a more pronounced it with the indices of solar activity easier and to use the monthly differential profile during the Maunder minimum (Eddy et al. and yearly differential rotation parameters from the Paper I. 1976). The time dependence of the solar sidereal differential ro- In the present paper small bright coronal structures (SBCS) tation velocity parameters A and B for the time interval − are used as tracers for the solar rotation determination and for 1998 2006 was analysed and represented graphically. Monthly a comparison of the deduced solar rotation parameters with in- and annual dependences of the solar rotation parameters A and B dices of solar activity on monthly and yearly temporal scales. are displayed in Figs. 1 and 2, respectively. Errors for the data The main physical characteristics of SBCS were summarised concerning monthly values are given in Paper I. by Brajša et al. (2004, 2008, and references therein), whilst On the time scale of months, the parameter A shows a slight their good latitudinal coverage on the Sun was described by, increase after the activity maximum, and the parameter B shows e.g. Brajša et al. (2005, and references therein). Solar rotation no evidence of a statistically significant time dependence (Figs. 1 determined tracing SBCS and its comparison with the side- and 2). On the annual time scale, the variation in parameter A real angular rotation velocity of subphotospheric layers were re- represented in Fig. 1 is more pronounced than those of the pa- cently analysed by Zaatri et al. (2009) and by Wöhl et al. (2010, rameter B given in Fig. 2. Solar sidereal equatorial rotation hereafter denoted as Paper I). In this work we use the same data velocity, represented by the parameter A, decreases up to the set and the same automatic method of data reduction for the de- year 2003 and then the A values increases (Fig. 1). termination of the monthly and yearly solar rotation parameters, as in Paper I. Besides the automatic method of data reduction ap- 3.2. A relationship between the solar rotation and activity plied in the present work, in Paper I also the interactive method was used, the results of the two methods were compared, the To examine correlation between the solar rotation and activity, north-south asymmetry of the solar rotation and the rigid rota- we performed a statistical analysis of the monthly and yearly tional component at high solar latitudes were investigated. Also, values of solar sidereal rotation velocity parameters A and B as a detailed comparison of the measured solar rotation of SBCS a function of different solar activity indices. Solar activity was with the results of other tracers and methods was presented in represented by the monthly relative sunspot number (SN), also Paper I. In the present paper we extend that analysis by studying called the Wolf number and monthly smoothed relative sunspot the temporal evolution of the solar rotation determined tracing number (SSN). This data series is provided by the solar influ- SBCS and a relationship between the solar rotation and activity ences data analysis center (SIDC) of the Royal Observatory of during most of the 23rd solar cycle (1998−2006). As proxies for Belgium (SIDC-team 1998−2006). The solar cycle dependence the solar activity we use the relative sunspot number (monthly, of solar sidereal rotation parameters Ast and Bst, obtained from

A17, page 2 of 6 R. Jurdana-Šepic´ et al.: On the solar rotation and activity in the period 1998–2006

Table 1. Statistical results for monthly and yearly solar sidereal rotation velocity parameters A and B as functions of solar activity.

Monthly ρ bMb Fig. t A = f (SN) –0.49 –0.0015 0.0003 3 −5.36 B = f (SN) –0.16 –0.0019 0.0012 4 −1.57 A = f (SSN) –0.57 –0.0018 0.0003 5 −6.57 B = f (SSN) –0.11 –0.0013 0.0013 6 −1.01 Ast = f (SN) –0.34 –0.0012 0.0003 −3.46 Bst = f (SN) –0.17 –0.0030 0.0018 −1.68 Ast = f (SSN) –0.40 –0.0015 0.0004 −4.22 Bst = f (SSN) –0.15 –0.0028 0.0020 −1.43 Yearly ρ bMb A = f (SN) –0.83 –0.0016 0.0004 3 −3.94 B = f (SN) –0.32 –0.0016 0.0018 4 −0.88 A = f (BIDV) –0.85 –0.0732 0.0172 7 −4.26 B = f (BIDV) –0.33 –0.0712 0.0774 8 −0.92

Notes. Functions of solar activity are represented by the monthly rela- tive sunspot number (SN), the monthly smoothed relative sunspot num- Fig. 1. Solar sidereal rotation parameter A as function of time; monthly ber (SSN), and the interplanetary magnetic field strength represented by (black dots) and yearly values (open circles) with corresponding error the IDV index. Ast and Bst denote solar sidereal rotation parameters ob- bars for yearly values and typical mean value uncertainties given in the tained from selected stable structures. ρ is the (Spearman) rank-order right corner of the figure. correlation coefficient, b the slope of the straight lines fitted through the set of data points under consideration, Mb is the standard error, and t is the test parameter for statistical significance of ρ.

Fig. 2. Solar sidereal rotation parameter B as function of time; monthly (black dots) and yearly values (open circles) with corresponding error bars for yearly values and typical uncertainties given in the right corner Fig. 3. Solar sidereal rotation parameter A as function of the relative of the figure. sunspot number (SN); monthly (black dots, full line) and yearly values (open circles, dotted line). Typical uncertainties are given in the right corner of the figure. selected stable structures, was also examined. The stable struc- tures were identified by searching for pairs of triplets with a difference in latitude of less than 0.5 deg and continuing val- In Figs. 7, 8 the error bars are not repeated, since they are the ues in longitude, i.e., in the central meridian distance (CMD). same as in previous cases for the yearly values. In general, less than half of the structures were found to be sta- According to results given in Table 1 and displayed in ble (Paper I). Figs. 3−8 we obtained the highest correlation between the so- The statistical analysis for the annual values of the solar side- lar sidereal rotation parameter A and the yearly activity proxies. real rotation parameters A and B have been determined as a func- Generally, solar sidereal rotation parameters obtained from se- tion of the solar activity (represented with SN and SSN)andalso lected stable structures Ast and Bst have lower correlations with as a function of the interplanetary magnetic field strength, de- solar activity than the parameters A and B. However, the param- rived from the IDV index (Svalgaard & Cliver 2005). The fol- eter Bst shows slightly higher correlation with solar activity in lowing values were calculated in the statistical analysis: the cor- comparison with the parameter B. Concerning yearly data, the relation coefficient ρ, the slope of the straight lines fitted through correlation of the parameter A with the relative sunspot num- the statistical set of data points b, and the corresponding standard ber (SN), and the interplanetary magnetic field strength derived error Mb. The values are given in Table 1 and represented in from the IDV index is very high, and in all cases correlation co- Figs. 3−8. Typical uncertainties are also given in these figures. efficient ρ exceeds 0.80.

A17, page 3 of 6 A&A 534, A17 (2011)

Fig. 7. Yearly values of the solar sidereal rotation parameter A as a func- Fig. 4. Solar sidereal rotation parameter B as function of the relative tion of the interplanetary magnetic field strength derived from the inter- sunspot number (SN); monthly (black dots, full line) and yearly values diurnal variability (IDV) index. (open circles, dotted line). Typical uncertainties are given in the right corner of the figure.

Fig. 5. Monthly values of the solar sidereal rotation parameter A as Fig. 8. Yearly values of the solar sidereal rotation parameter B as a func- a function of the monthly smoothed relative sunspot number (SSN). tion of interplanetary magnetic field strength derived from the interdi- Typical uncertainty is given in the right corner of the figure. urnal variability (IDV) index.

4. Discussion, interpretation, and conclusion In this paper we have found a dependence of the solar rotation on the phase of the solar cycle (Figs. 1, 2). This is clearly visible for the solar rotation parameter A, corresponding to the equato- rial rotation velocity (Fig. 1), whilst the results are not conclu- sive for the gradient of the solar differential rotation, expressed by parameter B (Fig. 2). As expected, these results are visible more clearly when the yearly values (Figs. 1, 2) are considered instead of the monthly values (Figs. 1, 2), where the noise is much greater. We note that Gissot (2009) developed an automatic proce- dure for identifying and tracking structures in sequences of full- disc solar images. This procedure is based on a tracking al- gorithm of solar features derived from wavelet transformation of the images. Among various subtopics, that procedure was applied to determining the solar rotation using the archive of Fig. 6. Monthly values of the solar sidereal rotation parameter B as EIT He II 30.4 nm images that now covers more than the 23rd so- a function of the monthly smoothed relative sunspot number (SSN). lar cycle. As in the present paper, stronger noise was found in Typical uncertainty is given in the right corner of the figure. the temporal evolution of the rotation parameter B than of the

A17, page 4 of 6 R. Jurdana-Šepic´ et al.: On the solar rotation and activity in the period 1998–2006 rotation parameter A. In spite of some similarities, the depen- We note that the equatorial rotation velocity of SBCS, aver- dence of the solar rotation parameter A on the time is not the aged over the whole observing period 1998−2006, is very sim- same for the two investigations (present work and Gissot 2009). ilar to the equatorial rotation velocity of sunspots and sunspot Furthermore, a relationship between the solar rotation and groups (solar rotation parameter A, Table 16 in Paper I), so the activity, expressed by different proxies (the monthly relative solar rotation parameter A determined tracing SBCS is a very sunspot number, the smoothed monthly relative sunspot num- good proxy for the solar equatorial rotation determined by ber, the yearly relative sunspot number, and the IDV index) was sunspots and sunspot groups. also investigated. These results are presented in Table 1 and The most probable theoretical interpretation of the observed in Figs. 3−8. Again, the statistically significant correlation was phenomena is based on the interaction between the solar dif- found for the solar rotation parameter A, whilst a very low and ferential rotation and magnetic field, as discussed in detail in insignificant correlation was obtained for rotation parameter B. the papers by Brajša et al. (2006, 2007, and references therein). As can be seen in Table 1, the slope of the linear least-square fit- A higher rotation velocity at low latitudes is consistent with a ting is statistically significant on the 3σ level for all cases where lower magnetic activity on the Sun. This is the result of the the parameter A is considered. For them, the correlation coeffi- numerical simulations performed by Brun (2004), and by Brun cient is in the range from 0.34 to 0.57 for the monthly values and et al. (2004), in agreement with an analytical approach to solve in the range from 0.83 to 0.85 for the yearly values (Table 1), the angular momentum equation (Lanza 2006, 2007). clearly indicating a higher correlation when the yearly values Finally, we can raise the question why in our analysis no sig- are considered. Statistical significance of correlations was deter- nificant dependence of the differential rotation parameter B on mined testing the null hypothesis (ρ = 0, “no correlation ex- time and/or on the activity level was found. During the periods ists”) at the level of 0.05. For monthly values the degree of free- of high magnetic activity on the Sun, a more rigid (i.e., a less dom is 91. As seen from Table 1 calculated test parameter t is differential) rotational component should be observed, yielding a for all “A-correlations” greater than Student’s distribution value lower absolute B value (Brun 2004; Lanza 2007). The most prob- of t at a chosen level of significance of say t(0.05, 91) = 1.64; able explanation for our lack of finding this is that to determine therefore, the null hypothesis is rejected. On the other hand, the solar rotation parameters we used data from all latitudes, all calculated values of t for “B-correlations” except for the even in polar regions, i.e., above 80 deg of latitude (Paper I). Bst = f (SN) dependence given in Table 1 are lower than the At those high latitudes the noise and various sources of system- value t(0.05, 91) = 1.64. Therefore for B values the null hypoth- atic errors are much larger than at the low latitudes, so these ff esis is accepted in all cases except for Bst = f (SN). For yearly e ects smear the variations of the parameter B, whilst having values, where the degree of freedom is 7 and t(0.05, 7) = 1.89 the barely any influence on the parameter A. Also, the effect of the null hypothesis is rejected for both A correlations and accepted height of the tracers on the measured rotation velocity increases for both B correlations (Press et al. 1989). with the latitude and has the strongest influence in polar regions Our most important result is that during the maximum of the (e.g., Roša et al. 1998). Moreover, the north-south rotational solar cycle 23 and just after it, the equatorial solar rotation ve- asymmetry in our data set is much greater at medium and high locity, expressed by the rotation parameter A,waslowerthan latitudes than at the lower ones (Paper I). As a suggestion for a in other phases of the cycle, when there was less activity. This future work, we will repeat the analysis by imposing different agrees with the high correlation between the rotation parame- cut-offs in the latitude (especially excluding the polar regions) ter A and various indices of solar activity. This result is consis- and performing the data reduction separately for the two solar tent with many other observational findings, obtained by differ- hemispheres. ent tracers and methods, so a higher average rotation velocity in activity minima was found by many authors who analysed Acknowledgements. This project was started with the support of the Alexander the solar rotation using various sunspot data series by different von Humboldt Foundation and is related to the SOHO-EIT Proposal methods (Lustig 1983; Gilman & Howard 1984; Balthasar et al. Brajsa_206: “an analysis of the solar rotation velocity by tracing coronal fea- 1986; Hathaway & Wilson 1990; Gupta et al. 1999; Khutsishvili tures” (http://umbra.nascom.nasa.gov/eit/proposals/) submitted in et al. 2002; Zuccarello & Zappalá 2003; Brajša et al. 2006). March 1999 by R. Brajša, B. Vršnak, V. Ruždjak, D. Roša, H. Wöhl, & F. ff Clette. SOHO is a project of international cooperation between ESA and NASA. Moreover, Chandra et al. (2010) measured the di erential rota- We would like to thank the EIT team for developing and operating the instru- tion of the soft X-ray solar corona and investigated its variation ment, and F. Clette, J.-F. Hochedez, and J. de Patoul for providing the EIT im- with the solar activity. They find that the equatorial rotation ve- ages and helpful discussions. The authors would also like to thank E. W. Cliver locity correlates well with the corresponding solar cycle phase, and J. Rybák for useful comments and suggestions. This work is sponsored by the Air Force Office of Scientific Research, Air Force Material Command, whilst parameters B and C do not show any systematic varia- USAF, under grant number FA8655-07-1-3093. Further, the research leading to tions. Our results obtained in the present paper agree with the the results presented in this paper has partly received funding from European results of Chandra et al. (2010). Community’s Seventh Framework Programme (FP7/2007−2013) under grant Another implication of the correlation between the solar ro- agreement No. 218816, and R.B. and A.H. acknowledge also the support from the Austrian-Croatian Bilateral Scientific Project, whilst S.F.G. acknowledges tation and activity is a higher solar rotation that should have been the support from the Belgian government through ESA-PRODEX. observed during the Maunder minimum. Some modern analyses of the data from the period of Maunder minimum suggest exactly this type of behaviour, although there is no general agreement on References this topic, as we have seen in Sect. 1. Since the floor of the inter- Abarbanell, C., & Wöhl, H. 1981, Sol. Phys., 70, 197 planetary magnetic field is about 4 nT, this is an expected value Altrock, R. C. 2003, Sol. Phys., 213, 23 for the Maunder minimum (Svalgaard & Cliver 2007, 2010), and Antia, H. M., & Basu, S. 2001, ApJ, 559, L67 we can find a corresponding value for the solar rotation param- Arévalo, M. J., Gomez, R., Vázquez, M., Balthasar, H., & Wöhl, H. 1982, A&A, eter A, describing the equatorial rotation velocity, from Fig. 7. 111, 266 / Balthasar, H., & Wöhl, H. 1980, A&A, 92, 111 It is about 0.2 deg day higher than average value for sunspots in Balthasar, H., Vázquez, M., & Wöhl, H. 1986, A&A, 155, 87 usual solar activity conditions and in good agreement with the Brajša, R., Vršnak, B., Ruždjak, V., et al. 1992, Hvar Obs. Bull., 16, 13 value reported by Eddy et al. (1976) for the Maunder minimum. Brajša, R., Ruždjak, V., Vršnak, B., et al. 1997, Sol. Phys., 171, 1

A17, page 5 of 6 A&A 534, A17 (2011)

Brajša, R., Ruždjak, V., Vršnak, B., et al. 1999, Sol. Phys., 184, 281 Lanza, A. F. 2006, MNRAS, 373, 819 Brajša, R., Ruždjak, V., Vršnak, B., et al. 2000, Sol. Phys., 196, 279 Lanza, A. F. 2007, A&A, 471, 1011 Brajša, R., Wöhl, H., Vršnak, B., et al. 2004, A&A, 414, 707 Lustig, G. 1983, A&A, 125, 355 Brajša, R., Wöhl, H., Vršnak, B., et al. 2005, Sol. Phys., 231, 29 Mendoza, B. 1999, Sol. Phys., 188, 237 Brajša, R., Ruždjak, D., & Wöhl, H. 2006, Sol. Phys., 237, 365 Meunier, N., Nesme-Ribes, E., & Grosso, N. 1997, A&A, 319, 673 Brajša, R., Wöhl, H., Ruždjak, D., et al. 2007, Astron. Nachr., 328, 1013 Nash, A. G., Sheeley, N. R. J., & Wang, Y.-M. 1988, Sol. Phys., 117, 359 Brajša, R., Wöhl, H., Vršnak, B., et al. 2008, CEAB, 32, 165 Press, W. H., Flannery, B. P., Teukolsky, S. A., & Vetterling, W. T. 1989, Brun, A. S. 2004, Sol. Phys., 220, 333 Numerical Recipes in Pascal (Cambridge University Press), 498 Brun, A. S., Miesch, M. S., & Toomre, J. 2004, ApJ, 614, 1073 Pulkkinen, P., & Tuominen, I. 1998, A&A, 332, 755 Casas, R., Vaquero, J. M., & Vázquez, M. 2006, Sol. Phys., 234, 379 Ribes, J. C., & Nesme-Ribes, E. 1993, A&A, 276, 549 Chandra, S., & Vats, H. O. 2011, MNRAS, 414, 3158 Roša, D., Vršnak, B., Božic,´ H., et al. 1998, Sol. Phys., 179, 237 Chandra, S., Vats, H. O., & Iyer, K. N. 2009, MNRAS, 400, L34 Rybák, J. 1994, Sol. Phys., 152, 161 Chandra, S., Vats, H. O., & Iyer, K. N. 2010, MNRAS, 407, 1108 Schröter, E.-H. 1985, Sol. Phys., 100, 141 Eddy, J. A., Gilman, P. A., & Trotter, D. E. 1976, Sol. Phys., 46, 3 Sheeley, N. R. J., Nash, A. G., & Wang, Y.-M. 1987, ApJ, 319, 481 Gigolashvili, M. S., Mdzinarishvili, T. G., Japaridze, D. R., & Chargeishvili, Sheeley, N. R. J., Wang, Y.-M., & Nash, A. G. 1992, ApJ, 401, 378 B. B. 2003, New Astron., 8, 529 SIDC-team 1998−2006, World Data Center for the Sunspot Index, Monthly Gilman, P., & Howard, R. 1984, ApJ, 283, 385 Report on the International Sunspot Number, Online catalogue of the Gissot, S. 2009, in Ph.D. Thesis, Université Catholique de Louvain, Ch. 5, 79 sunspot index, Royal Observatory of Belgium, http://www.sidc.be/ Gupta, S. S., Sivaraman, K. R., & Howard, R. F. 1999, Sol. Phys., 188, 225 sunspot-data Hathaway, D. H., & Wilson, R. M. 1990, ApJ, 357, 271 Snodgrass, H. 1992, in The Solar Cycle, ed. K. Harvey (San Francisco: ASP), Heristchi, D., & Mouradian, Z. 2009, A&A, 497, 835 ASP Conf. Ser., 27, 205 Howard, R. 1984, ARA&A, 22, 131 Stix, M. 2002, in The Sun (Berlin: Springer-Verlag), 2nd edn., 277 Howe, R., Christensen-Dalsgaard, J., Hill, F., et al. 2000a, Science, 287, 2456 Suzuki, M. 1998, Sol. Phys., 178, 259 Howe, R., Christensen-Dalsgaard, J., Hill, F., et al. 2000b, ApJ, 533, L163 Svalgaard, L., & Cliver, E. 2005, J. Geophys. Res., 110, A12103 Howe, R., Hill, F., Komm, R. W., et al. 2001, in Recent insights into the physics Svalgaard, L., & Cliver, E. 2007, ApJ, 661, L203 of the Sun and heliosphere: Highlights from SOHO and other space missions, Svalgaard, L., & Cliver, E. 2010, J. Geophys. Res., 115, A09111 ed. P. Brekke, B. Fleck, & J. Gurman (San Francisco: ASP), 40 Vaquero, J. M., Sánchez-Bajo, F., & Gallego, M. C. 2002, Sol. Phys., 207, 219 Japaridze, D. R., & Gigolashvili, M. S. 1992, Sol. Phys., 141, 267 Vats, H. O., Chandra, S., & Iyer, K. N. 2010, in Magnetic Coupling between the Javaraiah, J. 2003, Sol. Phys., 212, 23 Interior and Atmosphere of the Sun, ed. S. S. Hasan, & R. J. Rutten (Berlin Javaraiah, J., & Ulrich, R. K. 2006, Sol. Phys., 237, 245 Heidelberg: Springer), 526 Javaraiah, J., Bertello, L., & Ulrich, R. K. 2005, Sol. Phys., 232, 25 Wöhl, H., Brajša, R., Hanslmeier, A., & Gissot, S. 2010, A&A, 520, A29 Kambry, M. A., & Nishikawa, J. 1990, Sol. Phys., 126, 89 (Paper I) Khutsishvili, E. V., Gigolashvili, M. S., & Kvernadze, T. M. 2002, Sol. Phys., Zaatri, A., Wöhl, H., Roth, M., Corbard, T., & Brajša, R. 2009, A&A, 504, 589 206, 219 Zuccarello, F., & Zappalá, R. 2003, Astron. Nachr., 324, 464

A17, page 6 of 6 4.3 Paper III: A Relationship between the Solar Rotation and Activ- ity Analysed by Tracing Sunspot Groups

The following article is published in Solar Physics, Volume 292, Issue 179, pp. 1-11 (2017). It is reproduced by the permission of Springer. The .pdf document is available in the online version of this journal and hereafter. My own contribution to this work was 20%.

The main aims of this paper are presented in Sec. 1. The main conclusions of this paper are listed here. The sunspot positions published in the GPR, USAF/NOAA and DPD databases in the period 1874 to 2016 were used to calculate yearly values of the solar differential rotation parameters A and B. A relationship between the solar rotation (differential rotation parameters) and activity (yearly mean total sunspot numbers) was investigated, using the recently revised sunspot number. The main finding of this paper is that the Sun rotates more differentially at the min- imum than at the maximum of activity during the 1977 – 2016 period. This is in agree- ment with theoretical predictions of a reduced differential rotation in the presence of strong magnetic fields. Another important finding is an inverse correlation between equatorial rotation and solar activity, that was found by many authors before and is now corroborated. This is not so clear for the differential rotation parameter B, where no secular changes were found. Also, it was shown that the secular decrease of the equatorial rotation rate, that is accompanied with the increase in activity, stopped in the last part of the twentieth century when solar activity started to decrease. On examining the strength of solar maxima, it was noted that if a peak of the equatorial rotation velocity is observed during the minimum, then the next maximum is weaker than the preceding one.

55 Solar Phys (2017) 292:179 DOI 10.1007/s11207-017-1199-8

A Relationship Between the Solar Rotation and Activity Analysed by Tracing Sunspot Groups

Domagoj Ruždjak1 · Roman Brajša1 · Davor Sudar1 · Ivica Skokic´1 · Ivana Poljanciˇ cBeljan´ 2

Received: 30 August 2017 / Accepted: 27 October 2017 © Springer Science+Business Media B.V., part of Springer Nature 2017

Abstract The sunspot position published in the data bases of the Greenwich Photohelio- graphic Results (GPR), the US Air Force Solar Optical Observing Network and National Oceanic and Atmospheric Administration (USAF/NOAA), and of the Debrecen Photoheli- ographic Data (DPD) in the period 1874 to 2016 were used to calculate yearly values of the solar differential-rotation parameters A and B. These differential-rotation parameters were compared with the solar-activity level. We found that the Sun rotates more differentially at the minimum than at the maximum of activity during the epoch 1977 – 2016. An inverse cor- relation between equatorial rotation and solar activity was found using the recently revised sunspot number. The secular decrease of the equatorial rotation rate that accompanies the increase in activity stopped in the last part of the twentieth century. It was noted that when a significant peak in equatorial rotation velocity is observed during activity minimum, the next maximum is weaker than the previous one.

Keywords Differential rotation · Sunspot groups · Activity cycle

1. Introduction

Solar activity manifests itself in many diverse phenomena (sunspots, flares, coronal mass ejections, etc.) that vary on several distinct timescales. The most prominent manifestation of solar activity is the 11-year Schwabe cycle. Extensive reviews on the long-term behaviour of the solar activity and the Schwabe cycle are given by Hathaway (2010) and Usoskin (2017). It is generally accepted that the solar cycle is magnetic in nature and generated by dynamo processes within the Sun. The Sun’s magnetic field is maintained by its interaction with plasma motions, i.e. by the differential rotation of the convective zone. The physical processes underlying the complex magnetohydrodynamical system are still not fully under- stood, and a large number of dynamo models have been considered. An extensive review of

B D. Ruždjak [email protected]

1 Hvar Observatory, Faculty of Geodesy, University of Zagreb, Kaciˇ ceva´ 26, 10000 Zagreb, Croatia 2 Department of Physics, University of Rijeka, Radmile Matejciˇ c´ 2, 51000 Rijeka, Croatia 179 Page 2 of 11 D. Ruždjak et al. solar-dynamo models is given by Cameron, Dikpati, and Brandenburg (2016). Precise mea- surements of the solar differential rotation, its variations, and its correlation with the solar activity give important observational constraints on the modelling of the solar dynamo. The solar differential rotation has been determined by many authors using various meth- ods. Most often, sunspots and sunspot groups have been used as tracers because they are well-defined structures with sufficient lifetimes, and long series of observations are avail- able. Their disadvantages are their complex structure and evolution, limb-effects, and narrow latitudinal distribution. Most extensively, the Greenwich Photoheliographic Results (GPR) were used, e.g. by Newton and Nunn (1951), Balthasar, Vazquez, and Wöhl (1986), Bra- jša et al. (2002), and Ruždjak et al. (2005), among many others. The GPR data were often extended by data from the US Air Force Solar Optical Observing Network and National Oceanographic and Atmospheric Administration (USAF/NOAA) (Pulkkinen and Tuomi- nen, 1998; Zuccarello and Zappalá, 2003; Ruždjak et al., 2004; Brajša, Ruždjak, and Wöhl, 2006; Sudar et al., 2014; Javaraiah and Bertello, 2016). In addition to GPR and USAF/NOAA sunspot data, a number of other datasets were also frequently used, e.g. the Mount Wilson dataset (Howard, Gilman, and Gilman, 1984; Gilman and Howard, 1984), which consists of the areas and positions of the sunspots measured on the white-light pho- tographs of the Sun taken with the 30 cm diameter lens at the Tower Telescope at Mount Wilson Observatory, Kodaikanal data (Gupta, Sivaraman, and Howard, 1999), Kanzelhöhe data (Lustig, 1983; Poljanciˇ c´ Beljan et al., 2014; Poljanciˇ c-Beljan´ et al., 2017), and Debre- cen Photoheliographic Data (DPD: Sudar et al., 2017). In addition to sunspots and sunspot groups, other data were used for the determination of solar differential rotation: for instance, Doppler shifts (Howard, 1976), coronal holes (Nash, Sheeley, and Wang, 1988), faculae (Meunier, Nesme-Ribes, and Collin, 1997), Hα filaments (Brajša et al., 1997), global mag- netic fields (Obridko and Shelting, 2001), helioseismology (Howe et al., 2000; Antia and Basu, 2001; Komm, Howe, and Hill, 2017), solar radio emission (Li et al., 2012), coronal bright points (CBP) (Wöhl et al., 2010; Sudar et al., 2015, 2016), and coronal green-line emission (Rybak, 1994; Badalyan and Obridko, 2017). The inferred solar rotation was often analysed for long-term changes or changes within a cycle (see, e.g., Howard, 1976; Gilman and Howard, 1984; Gupta, Sivaraman, and Howard, 1999). More recently, Brajša, Ruždjak, and Wöhl (2006), using rotation-rate residuals calcu- lated from sunspot groups, found a secular deceleration of rotation and faster rotation at the minimum than at the maximum of the solar cycle. Jurdana-Šepic´ et al. (2011), using rotation rates obtained by tracing CBP, investigated the relationship between the solar-rotation pa- rameters and solar activity that is expressed by the sunspot number, and they found a signif- icant correlation for parameter A and a small, statistically insignificant, negative correlation for the differential-rotation parameter B.Liet al. (2014) revisited published solar-rotation data obtained from Hα charts and sunspots and confirmed the secular deceleration and neg- ative correlation between the differential-rotation parameter A and solar activity. Most re- cently, Badalyan and Obridko (2017) analysed the coronal green-line data from 1943 up to the present and found that the equatorial rotation rate increases in the epochs of minimum between the even- and odd-numbered solar cycles, and it reaches its minimum between the odd and even cycles, and that the differential gradient increases in the even cycles, with its highest values near the maximum of the activity cycle. On the other hand, theoretical calcu- lations (Brun, 2004; Brun, Miesch, and Toomre, 2004; Lanza, 2006, 2007) predict that the Sun rotates more rigidly at the maximum than at the minimum of activity. To gain insight into these discrepancies between observations and theoretical predictions, in this work we examine the relationship between solar activity and differential-rotation pa- rameters determined from the sunspot-group position data from GPR, USAF/NOAA, and Solar Rotation and Activity Page 3 of 11 179

DPD in the period from 1874 to 2016. The sunspot-group position data were used to calcu- late yearly differential-rotation parameters, and the obtained parameters were then compared with the solar activity. Furthermore, during 2011 – 2014, a series of four workshops was or- ganised with the goal of providing a community-vetted time series of sunspot numbers for use in long-term studies (Cliver, Clette, and Svalgaard, 2013;Cliveret al., 2015). This effort resulted in a recalibration of sunspot and group numbers (Clette et al., 2014, 2015, 2016). The revised sunspot numbers are used as a measure of solar activity in this work.

2. Data and Reduction

In this work the sunspot-group data from the GPR (1874 – 1976), USAF/NOAA Sunspot Data, and DPD (Baranyi, Gyori,˝ and Ludmány, 2016;Gyori,˝ Ludmány, and Baranyi, 2017) sunspot database (1977 – 2016) were used. Our own GPR digital dataset was used, while USAF/NOAA and DPD data were downloaded from the websites solarscience.msfc.nasa. gov/greenwch.shtm and fenyi.solarobs.csfk.mta.hu/DPD/, respectively. The given daily po- sitions of the sunspot groups were used to obtain the rotational velocities. For the groups for which the central meridian distance (CMD) was smaller than 58◦, which corresponds to about 0.85 of the projected solar radius (see Balthasar, Vazquez, and Wöhl, 1986), the rotation rate was calculated by dividing the CMD differences by the elapsed time. The ob- tained synodic rotation velocities were then transformed into sidereal ones by the procedure described by Roša et al. (1995), Brajša et al. (2002), and Skokic´ et al. (2014). Finally, the sidereal rotation velocities of less than 8◦ day−1 and/or exceeding 19◦ day−1 were regarded as erroneous and were not considered in the analysis. These constraints left 92,762 rotation velocities (out of 161,714 sunspot position data) obtained from GPR, and 58,219 rotation ve- locities (out of 106,981) and 48,814 (out of 84,449) for the DPD and USAF/NOAA sunspot databases, respectively. The solar differential rotation is usually represented by

ω = A + B sin2 ϕ + C sin4 ϕ, (1) where ω is the sidereal angular rotation velocity, ϕ is the heliographic latitude, and A, B,and C are differential-rotation parameters. Because of the latitudinal distribution of sunspots, the last term can be neglected when using sunspot motions to derive the parameters, i.e. C = 0. Differential-rotation parameters A and B were obtained by fitting the resulting equation to all points in the dataset. The results for individual datasets are presented in Table 1.The parameters A and B were also calculated separately for each year in the time span of the datasets (1874 – 2016). Data from both of the solar hemispheres were treated together to give a statistically significant number of data points for the determination of differential-rotation coefficients in the years near minimum of activity.

Table 1 Differential-rotation ◦ −1 ◦ −1 parameters for various datasets. Time span A [ day ] B [ day ] Dataset 1: GPR; 2: USAF/NOAA; 3: DPD; 4: GPR+USAF/NOAA; 1874 – 1976 14.528 ± 0.006 −2.77 ± 0.05 1 + and 5: GPR DPD. 1977 – 2016* 14.44 ± 0.01 −2.54 ± 0.08 2 1977 – 2016 14.403 ± 0.009 −2.44 ± 0.08 3 1874 – 2016* 14.501 ± 0.005 −2.71 ± 0.05 4 *The USAF/NOAA set contains 1874 – 2016 14.483 ± 0.005 −2.67 ± 0.05 5 only Jan.–Sep. data for 2016. 179 Page 4 of 11 D. Ruždjak et al.

The solar-activity data (total sunspot number) were taken from the WDC-SILSO (sidc.oma.be/silso/) of the Royal Observatory of Belgium, Brussels. We used version 2.0 of the data, which contains a new, entirely revised data series and has been available since 1 July 2015. The yearly mean total sunspot number, obtained by taking a simple arithmetic mean of the daily total sunspot number over all of the days of one year, was taken as the measure for the solar activity of a given year.

3. Results

3.1. Secular Changes in Solar Rotation and Activity

For illustration of the secular changes of solar rotation and activity, we present in Figure 1 the yearly values of the differential-rotation coefficient A (equatorial rotation velocity) to- gether with the yearly mean total sunspot number. Rotation rates are denoted with circles and mean sunspot numbers with asterisks. The lines represent the linear fits to the data in the period 1880 to 2016. Only the GPR and DPD rotation data are shown. During the last 150 years, the solar activity was growing on average, reaching its maximum in the second half of the twentieth century, and it has been decreasing since. This can be noted by examining the amplitude (strength) of the maxima of Solar Cycles 12 – 24, which are represented by the yearly mean total sunspot numbers in the lower part of Figure 1. This increase/decrease would be monotonic, but the Cycle 14, 16, and 20 maxima are too weak to support this in- terpretation. On the other hand, the solar equatorial rotation shows a steady decrease during this period. The exceptions are the data for the first six years (1874 – 1879), whose values (except 1878) are systematically lower than the rest of the data. These data (circled points in the left part of Figure 1) were excluded from the calculation of the secular trend. The linear fits give for the increase in activity 0.2±0.1 year−1 and −0.0014±0.0003◦ day−1 year−1 be- tween 1880 and 2016 and −0.0012 ± 0.0003◦ day−1 year−1 for 1880 – 2013 for the decrease in equatorial rotation rate. The obtained value is the same as the values obtained by Brajša, Ruždjak, and Wöhl (2006) for rotation rate residuals and by Li et al. (2014) for comparison of previous results for different methods and datasets.

Figure 1 Secular variations of solar rotation and activity. The yearly values of the equatorial rotation rate A (black circles) are shown together with the yearly mean total sunspot number (asterisks). The lines represent linear fits of the data for the time interval 1880 – 2016. The GPR and DPD data are shown. The circled points denote the suspiciously low values of the equatorial rotation at the edges of the observing interval. Solar Rotation and Activity Page 5 of 11 179

It is worth pointing out that the equatorial rotation-velocity minimum in 2014 – 2016 (circled points in the right part of Figure 1) is not present in the USAF/NOAA data. The USAF/NOAA data give a rise of the equatorial rotation of 0.004 ± 0.002◦ day−1 year−1 for the 1977 – 2016 interval, but they have suspiciously low values of the coefficient A for the first two years (1977 – 1978). These results show how difficult it can be to determine the sec- ular trend when it is “obscured” by the cycle-related variations. It is unfortunate that all of the datasets have some extreme points near the edges. In addition to the suspiciously low val- ues at the beginning, the GPR dataset has its maximum yearly value of 14.90 ± 0.15◦ day−1 near the end of observing period (1965) because of the cycle-related variations, which also complicates the determination of the secular trend. To work around this problem, besides omitting the suspicious points, all further analyses are made by taking into account the errors of yearly values of solar differential-rotation parameters. The results are similar if the “problematic” points are omitted or all data are considered, but with errors taken into account. A similar deceleration is obtained (−0.0016 ± 0.0002◦ day−1 year−1) for 1874 – 2016 when the errors are considered. Individual sets yield −0.0007 ± 0.0003◦ day−1 year−1 (GPR 1874 – 1976), −0.0007 ± 0.0010◦ day−1 year−1 (DPD 1977 – 2016), and 0.004 ± 0.001◦ day−1 year−1 (USAF/NOAA 1977 – 2016). These results indicate that the secular de- crease of the rotation velocity stopped sometime in the second half of the twentieth century. It was mentioned that the maximum equatorial rotation rate was observed in 1965. The peak of equatorial rotation is observed during the minimum of . The (weighted) mean value of the equatorial rotation during the minimum (14.69 ± 0.05, 1963 – 65) is significantly (2σ ) higher than the value for the preceding and following years (14.47± 0.03, 1960 – 62 and 14.52 ± 0.03, 1966 – 68). A similar value of 14.88 ± 0.18◦ day−1 was observed during the minimum of Solar Cycle 14. It can be noted in Figure 1 that Solar Cy- cles 14 and 20 have a smaller amplitude than “expected” (smaller than what is needed for the envelope to be monotonic). Similar peaks of rotation velocity are visible for Cycles 23 and 24, but the peak is not observed for Cycle 16 (although there is a jump of ≈ 0.25◦ day−1 during the previous minimum). Therefore, it can be concluded that if the peak of the equa- torial rotation velocity is observed during the minimum, the following maximum is weaker than the previous one. The secular variation of the differential-rotation parameter B is shown in the upper panel of Figure 2. From inspecting the figure, it is hard to see any correlation between the changes in so- lar activity and differential-rotation parameter B. During the maximum of activity, B has values similar to its average value, while it oscillates during the solar-activity minima. This behaviour is due to the small amount of data in years near the minimum. This is illustrated in the lower panel of the figure where we show the total number of data. It can be seen that in some solar minima, this number is lower than 50 (dotted-horizontal line in the figure). In Table 2 the values for differential-rotation parameters for several minima and maxima are presented. From inspecting the values in the table, two conclusions can be drawn. First, the values of equatorial rotation velocity are on average higher at the minimum than at the maximum of activity (weighted means are 14.57 ± 0.04◦ day−1 and 14.44 ± 0.02◦ day−1 for the minimum and maximum of activity, respectively, which is above the 2σ level). Second, about 50% of the values for the differential-rotation parameter B in the years near minimum of activity have large errors. In the analysis of such a dataset, it is vital that the errors of individual values are considered. Finally, we would like to draw attention to the differential-rotation parameters for in- dividual sets presented in Table 1. The equatorial rotation velocity from the GPR is sig- nificantly higher than the one in USAF/NOAA and DPD data. The difference between 179 Page 6 of 11 D. Ruždjak et al.

Figure 2 Secular variations of the differential-rotation parameter B (upper panel). GPR and DPD data are represented by points connected with the dashed line, while USAF/NOAA data are shown with crosses con- nected by the dotted line.Inthelower panel we show the number of data available for a given year. GPR and DPD data are represented with the solid line and USAF/NOAA with the dashed line.Thehorizontal-dotted line represents N = 50 level. The yearly mean total sunspot number is represented by the grey line in both panels.

◦ − Table 2 Values and errors of differential-rotation coefficients [ day 1] for several solar-cycle minimum and maximumyearsthatarecoveredbythedata(1:GPR;2:USAF/NOAA;and3:DPD).

Minimum Maximum Year ABNYear ABN

1878 14.80 ± 0.39 −20.0 ± 19 43 1884 14.27 ± 0.07 −1.7 ± 1.2 1138 1 1890 14.65 ± 0.24 −3.2 ± 1.5 158 1893 14.67 ± 0.04 −2.8 ± 0.4 1817 1 1902 14.66 ± 0.29 −0.2 ± 2.2 68 1906 14.58 ± 0.06 −3.1 ± 0.8 1096 1 1986 14.55 ± 0.10 −2.5 ± 0.9 226 1989 14.38 ± 0.05 −2.0 ± 0.4 2602 2 1986 14.55 ± 0.08 −3.4 ± 0.8 297 1989 14.38 ± 0.04 −2.6 ± 0.3 3685 3 1996 14.47 ± 0.14 0.3 ± 1.6 158 2000 14.49 ± 0.05 −2.8 ± 0.4 2113 2 1996 14.70 ± 0.09 −2.6 ± 1.2 238 2000 14.37 ± 0.04 −2.7 ± 0.3 3658 3 2008 14.63 ± 0.27 −1.0 ± 1.9 66 2014 14.46 ± 0.05 −1.7 ± 0.7 1467 2 2008 14.36 ± 0.15 0.6 ± 1.1 137 2014 14.26 ± 0.05 −3.7 ± 0.6 3344 3

USAF/NOAA and DPD is (just 0.001◦ day−1) inside the 2σ range. The absolute value of parameter B for the GPR data is 2σ larger than the value for DPD and 1.77σ larger than the value for the USAF/NOAA, while DPD and USAF/NOAA have values within 1σ .This means that the equatorial rotation rate is on average higher during 1874 – 1976 than in the period 1977 – 2016, and maybe the Sun rotated more differentially in the earlier epoch.

3.2. Correlation Between Solar Activity and Rotation

In Figure 3 we present the dependence of the solar differential-rotation parameters on the yearly mean total sunspot number. We show the USAF/NOAA dataset, which has the highest Solar Rotation and Activity Page 7 of 11 179

Figure 3 Differential-rotation parameters A and B for USAF/NOAA dataset shown versus yearly mean total sunspot number. The lines represent least-squares fits using the Marquardt–Levenberg algorithm.

Table 3 Least-squares fits of yearly values of solar differential-rotation parameters vs. yearly mean total sunspot numbers (YMTSN). (1: GPR; 2: USAF/NOAA; 3: DPD; 4: GPR+USAF/NOAA; and 5: GPR+ DPD).

− ◦ − − ◦ − Dataset Slope A (YMTSN) [10 3 day 1] Slope B (YMTSN) [10 3 day 1]

1 −0.1 ± 0.1 −1.6 ± 0.9 2 −0.6 ± 0.23.0 ± 1.5 3 −0.2 ± 0.23.4 ± 1.7 4 −0.3 ± 0.1 −0.6 ± 0.8 5 −0.2 ± 0.1 −0.2 ± 0.8

correlation. Solid lines represent the least-squares fits. To account for the errors of the solar differential-rotation parameters, the data were fitted using an implementation of the least- squares Marquardt–Levenberg algorithm. The results of fits for the different datasets are summarised in Table 3. As expected, the most significant values are obtained for the set with the highest correlation. The results, although not statistically significant for all datasets, show a negative correlation between the solar equatorial rotation rate and solar activity. The results for the differential-rotation parameter B are different for the various datasets. The GPR data result in an insignificant negative trend, while the data from USAF/NOAA and DPD result in a positive correlation with a 2σ statistical significance. Note that the values of B are negative, so the positive correlation means a lower absolute value during the maximum of activity. The results could imply that the correlation of the differential- rotation parameter B has changed. In the period 1977 – 2016, there is a positive correlation between the differential-rotation parameter B (significant on the 2σ level), and there is an (insignificant) negative correlation during 1874 – 1976. When the sets are combined, the larger GPR dataset prevails. 179 Page 8 of 11 D. Ruždjak et al.

4. Discussion and Conclusion

Our results show that the Sun rotates faster at low latitudes and more differentially at the minimum than at the maximum of activity. This result is in agreement with the result of Eddy, Gilman, and Trotter (1976), who measured the solar surface rotation from sunspot drawings during the Maunder minimum and found that solar equatorial rotation was 3 – 4% faster than today and that the differential rotation between the Equator and latitudes of 20◦ was about three times higher. However, Abarbanell and Wöhl (1981), using two original prints of Hevelius’ book, while Eddy, Gilman, and Trotter (1976) used a copy, found that the solar rotation was at the same level as today. The gradient of the differential rotation was found to be slightly steeper and not significantly different. More pronounced differ- ential rotation during the Maunder minimum was also found by Ribes and Nesme-Ribes (1993). Here, both secular and cycle-related changes of the equatorial rotation rate, which are anticorrelated with solar activity, were found. This is not so clear for the differential- rotation parameter B, where no secular changes were found. The higher average rotation velocities in the minimum compared to other phases of cycle, i.e. the inverse correlation be- tween solar equatorial-rotation rate and activity, was found by many authors using different methods and data (Lustig, 1983; Gilman and Howard, 1984; Balthasar, Vazquez, and Wöhl, 1986; Gupta, Sivaraman, and Howard, 1999; Zuccarello and Zappalá, 2003; Brajša, Ružd- jak, and Wöhl, 2006; Jurdana-Šepic´ et al., 2011;Liet al., 2014; Badalyan and Obridko, 2017). Brun (2004) numerically modelled the interaction of convection and rotation with the magnetic field in deep spherical shells and found that an increase in magnetic field would result in the reduction of differential rotation. The reduction of differential rotation is due to the braking effect exerted by non-axisymmetric magnetic fields. This means that in the pres- ence of strong magnetic fields, the net transport of angular momentum towards the Equator is less efficient. A similar result was obtained by Lanza (2006, 2007) by analytically solving the angular momentum transport equation within the convection zone of a rotating star. This explains the observed effect. At the minimum of activity, when the magnetic field is weaker, the angular momentum is transported more efficiently towards the Equator, which results in the observed increased equatorial rotation and in more pronounced differential rotation at the minimum of activity. The changes in the differential-rotation profile (differential-rotation parameter B)are much harder to detect than the changes in equatorial rotation. Jurdana-Šepic´ et al. (2011) analysed the relationship between solar activity and differential-rotation parameters ob- tained by tracing CBPs. They found an inverse correlation of the equatorial rotation rate with a several times larger slope than obtained here and no significant correlation between the parameter B and activity. They attributed the lack of correlation for parameter B to the errors of their data, which are more pronounced at high latitudes. A reduction of differential rotation in this work is detected (at the 2σ level) in the period 1977 – 2016, but not in the epoch 1874 – 1976. This is most probably due to the errors, which are more pronounced at the beginning of the dataset, where the amount of available data in years near minima is low. Badalyan and Obridko (2017) examined the 22-year cycle of differential rotation and the rule of Genevyshev–Ohl by using the green coronal line brightness data from 1943 on- ward and obtained a result consistent with an inverse correlation of the differential-rotation parameters and activity (see Figures 3 and 4 in Badalyan and Obridko, 2017). Their finding of larger increases in equatorial rotation velocity between even and odd cycles than between odd and even cycles is not visible in our data (see Figure 1). They did not find the difference between even- and odd-cycle variation of the differential-rotation parameter B,andtheir Solar Rotation and Activity Page 9 of 11 179 results show that the differential rotation is strongest just before maximum and weakest somewhere in between the minimum and maximum of activity. A similar result, although with much smaller statistical significance, is obtained in the GPR data. However, the ampli- tude and phase of the curve both depend on the sampling, i.e. how the phase of the cycle was chosen. On examining the strength of solar maxima, it was noted that if a peak of the equatorial rotation velocity is observed during the minimum, then the next maximum is weaker than the preceding one. The solar activity is high for 1950 – 2000, which can be regarded as the modern grand maximum of solar activity (Usoskin, 2017). If the magnetic energy were to exceed about 20% of the total kinetic energy, Maxwell stresses and magnetic torques may become strong enough to suppress the differential rotation almost entirely (Brun, Miesch, and Toomre, 2004, and references therein). This is not observed, therefore the Sun must have ways of avoiding this by expelling some of its magnetic flux. The observed stronger peaks of the equatorial rotational velocity might be a signature of such a process. The most significant such event is observed during the minimum following the maximum of , which was the strongest one in the studied epoch. Cycle 20 is much weaker than the three previous and three following cycles, as if magnetic energy were expelled or consumed. No significant change in the differential-rotation parameter B is observed to accompany the event. The main result of our investigation is the finding that the Sun rotates more differentially at the minimum than at the maximum of activity during the epoch 1977 – 2016. This is in agreement with theoretical predictions of a reduced differential rotation in the presence of strong magnetic fields. An inverse correlation between equatorial rotation and solar activity was found by many authors before and is corroborated here regardless of the recent revision of sunspot number. The secular decrease in rotation velocity that accompanies the increase in activity stopped in the last part of the twentieth century when the solar activity started to decrease. It was noted that when a significant peak of the equatorial rotation velocity is observed during the minimum of activity, the next maximum is weaker than the previous one. It was suggested that this finding might be connected to a decrease in magnetic energy of the Sun.

Acknowledgements The authors wish to thank Hubertus Wöhl for useful sugestions and a careful reading of the manuscript. We acknowledge the staff of the Royal Observatory of Belgium, Brussels, and the staff of the Heliophysical Observatory, Debrecen, Hungary for maintaining and organizing the WDC-SILSO and DPD databases, respectively. This work was partly supported by the Croatian Science Foundation under the project 6212 “Solar and Stellar Variability”, and in part by the University of Rijeka under project number 13.12.1.3.03.

Disclosure of Potential Conflicts of Interest The authors declare that they have no conflicts of interest.

References

Abarbanell, C., Wöhl, H.: 1981, Solar rotation velocity as determined from sunspot drawings of Hevelius, J. in the 17th-century. Solar Phys. 70, 197. DOI. ADS. Antia, H.M., Basu, S.: 2001, Temporal variations of the solar rotation rate at high latitudes. Astrophys. J. Lett. 559, L67. DOI. ADS. Badalyan, O.G., Obridko, V.N.: 2017, 22-year cycle of differential rotation of the solar corona and the rule by Gnevyshev–Ohl. Mon. Not. Roy. Astron. Soc. 466, 4535. DOI. ADS. Balthasar, H., Vazquez, M., Wöhl, H.: 1986, Differential rotation of sunspot groups in the period from 1874 through 1976 and changes of the rotation velocity within the solar cycle. Astron. Astrophys. 155, 87. ADS. 179 Page 10 of 11 D. Ruždjak et al.

Baranyi, T., Gyori,˝ L., Ludmány, A.: 2016, On-line tools for solar data compiled at the Debrecen observatory and their extensions with the Greenwich sunspot data. Solar Phys. 291, 3081. DOI. ADS. Brajša, R., Ruždjak, D., Wöhl, H.: 2006, Temporal variations of the solar rotation determined by sunspot groups. Solar Phys. 237, 365. DOI. ADS. Brajša, R., Ruždjak, V., Vršnak, B., Pohjolainen, S., Urpo, S., Schroll, A., Wöhl, H.: 1997, On the possible changes of the solar differential rotation during the activity cycle determined using microwave low- brightness regions and H filaments as tracers. Solar Phys. 171,1.DOI. ADS. Brajša, R., Wöhl, H., Vršnak, B., Ruždjak, D., Sudar, D., Roša, D., Hržina, D.: 2002, Differential rotation of stable recurrent sunspot groups. Solar Phys. 206, 229. DOI. ADS. Brun, A.S.: 2004, On the interaction between differential rotation and magnetic fields in the Sun. Solar Phys. 220, 333. DOI. ADS. Brun, A.S., Miesch, M.S., Toomre, J.: 2004, Global-scale turbulent convection and magnetic dynamo action in the solar envelope. Astrophys. J. 614, 1073. DOI. ADS. Cameron, R.H., Dikpati, M., Brandenburg, A.: 2016, The global solar dynamo. Space Sci. Rev.. DOI. ADS. Clette, F., Svalgaard, L., Vaquero, J.M., Cliver, E.W.: 2014, Revisiting the sunspot number. A 400-year per- spective on the solar cycle. Space Sci. Rev. 186, 35. DOI. ADS. Clette, F., Svalgaard, L., Vaquero, J.M., Cliver, E.W.: 2015, Revisiting the sunspot number. In: Balogh, A., Hudson, H., Petrovay, K., von Steiger, R. (eds.) The Solar Activity Cycle, Space Sci. Ser. ISSI 53, Springer, New York 35. DOI. ADS. Clette, F., Cliver, E.W., Lefèvre, L., Svalgaard, L., Vaquero, J.M., Leibacher, J.W.: 2016, Preface to topical issue: Recalibration of the sunspot number. Solar Phys. 291, 2479. DOI. ADS. Cliver, E.W., Clette, F., Svalgaard, L.: 2013, Recalibrating the sunspot number (SSN): The SSN workshops. Cent. Eur. Astrophys. Bull. 37, 401. ADS. Cliver, E.W., Clette, F., Svalgaard, L., Vaquero, J.M.: 2015, Recalibrating the sunspot number (SN): The 3rd and 4th SN workshops. Cent. Eur. Astrophys. Bull. 39,1.ADS. Eddy, J.A., Gilman, P.A., Trotter, D.E.: 1976, Solar rotation during the Maunder minimum. Solar Phys. 46, 3. DOI. ADS. Gilman, P.A., Howard, R.: 1984, Variations in solar rotation with the sunspot cycle. Astrophys. J. 283, 385. DOI. ADS. Gupta, S.S., Sivaraman, K.R., Howard, R.F.: 1999, Measurement of Kodaikanal white-light images, III: Ro- tation rates and activity cycle variations. Solar Phys. 188, 225. DOI. ADS. Gyori,˝ L., Ludmány, A., Baranyi, T.: 2017, Comparative analysis of Debrecen sunspot catalogues. Mon. Not. Roy. Astron. Soc. 465, 1259. DOI. ADS. Hathaway, D.H.: 2010, The solar cycle. Liv. Rev. Solar Phys. 7,1.DOI. ADS. Howard, R.: 1976, A possible variation of the solar rotation with the activity cycle. Astrophys. J. Lett. 210, L159. DOI. ADS. Howard, R., Gilman, P.I., Gilman, P.A.: 1984, Rotation of the sun measured from Mount Wilson white-light images. Astrophys. J. 283, 373. DOI. ADS. Howe, R., Christensen-Dalsgaard, J., Hill, F., Komm, R.W., Larsen, R.M., Schou, J., Thompson, M.J., Toomre, J.: 2000, Dynamic variations at the base of the solar convection zone. Science 287, 2456. DOI. ADS. Javaraiah, J., Bertello, L.: 2016, Short-term variations in the equatorial rotation rate of sunspot groups. Solar Phys. 291, 3485. DOI. ADS. Jurdana-Šepic,´ R., Brajša, R., Wöhl, H., Hanslmeier, A., Poljanciˇ c,´ I., Svalgaard, L., Gissot, S.F.: 2011, A relationship between the solar rotation and activity in the period 1998 – 2006 analysed by tracing small bright coronal structures in SOHO-EIT images. Astron. Astrophys. 534, A17. DOI. ADS. Komm, R., Howe, R., Hill, F.: 2017, Solar-cycle variation of subsurface-flow divergence: A proxy of magnetic activity? Solar Phys. 292, 122. DOI. ADS. Lanza, A.F.: 2006, On the time dependence of differential rotation in young late-type stars. Mon. Not. Roy. Astron. Soc. 373, 819. DOI. ADS. Lanza, A.F.: 2007, Angular momentum conservation and torsional oscillations in the Sun and solar-like stars. Astron. Astrophys. 471, 1011. DOI. ADS. Li, K.J., Shi, X.J., Feng, W., Xie, J.L., Gao, P.X., Zhan, L.S., Liang, H.F.: 2012, Long-term variations of the coronal rotation and solar activity. Mon. Not. Roy. Astron. Soc. 423, 3584. DOI. ADS. Li, K.J., Feng, W., Shi, X.J., Xie, J.L., Gao, P.X., Liang, H.F.: 2014, Long-term variations of solar differential rotation and sunspot activity: Revisited. Solar Phys. 289, 759. DOI. ADS. Lustig, G.: 1983, Solar rotation 1947 – 1981 – determined from sunspot data. Astron. Astrophys. 125, 355. ADS. Meunier, N., Nesme-Ribes, E., Collin, B.: 1997, Cycle 19 facula dynamics, II: Meridional circulation, RMS velocity, and Reynolds stresses. Astron. Astrophys. 319, 683. ADS. Solar Rotation and Activity Page 11 of 11 179

Nash, A.G., Sheeley, N.R. Jr., Wang, Y.-M.: 1988, Mechanisms for the rigid rotation of coronal holes. Solar Phys. 117, 359. DOI. ADS. Newton, H.W., Nunn, M.L.: 1951, The Sun’s rotation derived from sunspots 1934 – 1944 and additional results. Mon. Not. Roy. Astron. Soc. 111, 413. ADS. Obridko, V.N., Shelting, B.D.: 2001, Rotation characteristics of large-scale solar magnetic fields. Solar Phys. 201,1.DOI. ADS. Poljanciˇ c´ Beljan, I., Jurdana-Šepic,´ R., Cargonja,ˇ M., Brajša, R., Hržina, D., Pötzi, W., Hanslmeier, A.: 2014, An analysis of the solar differential rotation from the Kanzelhöhe sunspot drawings. Cent. Eur. Astro- phys. Bull. 38, 87. ADS. Poljanciˇ c-Beljan,´ I., Jurdana-Šepic,´ R., Brajša, R., Sudar, D., Ruždjak, D., Hržina, D., Pötzi, W., Hanslmeier, A., Veronig, A., Skokic,´ I., Wöhl, H.: 2017, Solar diferential rotation in the period 1964 – 2016 deter- mined by the Kanzelhöhe data set. Astron. Astrophys. 606, A72. DOI. ADS. Pulkkinen, P., Tuominen, I.: 1998, Velocity structures from sunspot statistics in cycles 10 to 22, II: Latitudinal velocity and correlation functions. Astron. Astrophys. 332, 755. ADS. Ribes, J.C., Nesme-Ribes, E.: 1993, The solar sunspot cycle in the Maunder minimum AD1645 to AD1715. Astron. Astrophys. 276, 549. ADS. Roša, D., Brajša, R., Vršnak, B., Wöhl, H.: 1995, The relation between the synodic and sidereal rotation period of the Sun. Solar Phys. 159, 393. DOI. ADS. Ruždjak, D., Ruždjak, V., Brajša, R., Wöhl, H.: 2004, Deceleration of the rotational velocities of sunspot groups during their evolution. Solar Phys. 221, 225. DOI. ADS. Ruždjak, D., Brajša, R., Sudar, D., Wöhl, H.: 2005, The influence of the evolution of sunspot groups on the determination of the solar velocity field. Solar Phys. 229, 35. DOI. ADS. Rybak, J.: 1994, Rotational characteristics of the green solar corona: 1964 – 1989. Solar Phys. 152, 161. DOI. ADS. Skokic,´ I., Brajša, R., Roša, D., Hržina, D., Wöhl, H.: 2014, Validity of the relations between the synodic and sidereal rotation velocities of the Sun. Solar Phys. 289, 1471. DOI. ADS. Sudar, D., Skokic,´ I., Ruždjak, D., Brajša, R., Wöhl, H.: 2014, Tracing sunspot groups to determine angular momentum transfer on the Sun. Mon. Not. Roy. Astron. Soc. 439, 2377. DOI. ADS. Sudar, D., Skokic,´ I., Brajša, R., Saar, S.H.: 2015, Steps towards a high precision solar rotation profile: Results from SDO/AIA coronal bright point data. Astron. Astrophys. 575, A63. DOI. ADS. Sudar, D., Saar, S.H., Skokic,´ I., Poljanciˇ c´ Beljan, I., Brajša, R.: 2016, Meridional motions and Reynolds stress from SDO/AIA coronal bright points data. Astron. Astrophys. 587, A29. DOI. ADS. Sudar, D., Brajša, R., Skokic,´ I., Poljanciˇ c´ Beljan, I., Wöhl, H.: 2017, Meridional motion and Reynolds stress from Debrecen photoheliographic data. Solar Phys. 292, 86. DOI. ADS. Usoskin, I.G.: 2017, A history of solar activity over millennia. Liv. Rev. Solar Phys. 14,3.DOI. ADS. Wöhl, H., Brajša, R., Hanslmeier, A., Gissot, S.F.: 2010, A precise measurement of the solar differential rotation by tracing small bright coronal structures in SOHO-EIT images. Results and comparisons for the period 1998 – 2006. Astron. Astrophys. 520, A29. DOI. ADS. Zuccarello, F., Zappalá, R.A.: 2003, Angular velocity during the cycle deduced using the sunspot group age selection methodology. Astron. Nachr. 324, 464. DOI. ADS. 4.4 Paper IV: Meridional motions and Reynolds stress from SDO/AIA coronal bright points data

The following article is published in Astronomy and Astrophysics, Volume 587, Issue A29, pp. 1-6 (2016). It is reproduced by the permission of ESO. The .pdf document is available in the online version of this journal and hereafter. My own contribution to this work was 20%.

The main aims of this paper are presented in Sec. 1. The main conclusions of this paper are listed here. A combination of high resolution and high cadence of the SDO/AIA images, enabled a determination of very accurate velocity flows on the solar surface - differential rotation profile, meridional flows, torsional oscillations, and horizontal Reynolds stress. The accuracy of the calculated rotation enables the possibility to measure changes in the profile during the solar cycle. The meridional velocity is almost always positive, meaning the motions towards the poles, for all latitudes. Predominantly poleward flow was not found in other works dealing with tracers, but has been detected by using the Doppler method. Rotation velocity residuals show positive values for the latitude range 10◦-20◦, as well as for the latitudes above 50◦, which means that they show signs of torsional oscil- lations, e.g. this might correspond to a branches of torsional oscillation pattern towards the equator. This type of result is again more consistent with Doppler measurements than with sunspot data. Horizontal Reynolds stress was found to be smaller than in similar works, maybe due to the particular phase of the solar cycle 24, but still showed transfer of angular momentum towards the solar equator.

67 A&A 587, A29 (2016) Astronomy DOI: 10.1051/0004-6361/201527217 & c ESO 2016 Astrophysics

Meridional motions and Reynolds stress from SDO/AIA coronal bright points data

D. Sudar1,S.H.Saar2, I. Skokic´3, I. Poljanciˇ c´ Beljan4, and R. Brajša1

1 Hvar Observatory, Faculty of Geodesy, Kaciˇ ceva´ 26, University of Zagreb, 10000 Zagreb, Croatia e-mail: [email protected] 2 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA 3 Astronomical Institute of the Czech Academy of Sciences, Fricovaˇ 298, 251 65 Ondrejov,ˇ Czech Republic 4 Department of Physics, University of Rijeka, Radmile Matejciˇ c´ 2, 51000 Rijeka, Croatia Received 19 August 2015 / Accepted 16 December 2015

ABSTRACT

Context. It is possible to detect and track coronal bright points (CBPs) in Solar Dynamics Observatory/Atmospheric Imaging Assembly (SDO/AIA) images. A combination of high resolution and high cadence provides a wealth of data that can be used to determine velocity flows on the solar surface with very high accuracy. Aims. We derived a very accurate solar rotation profile and investigated meridional flows, torsional oscillations, and horizontal Reynolds stress based on ≈6 months of SDO/AIA data. Methods. We used a segmentation algorithm to detect CBPs in SDO/AIA images. We also used invariance of the solar rotation profile with central meridian distance (CMD) to determine the height of CBPs in the 19.3 nm channel. Results. The best fit solar rotation profile is given by ω(b) = (14.4060 ± 0.0051 + (−1.662 ± 0.050) sin2 b + (−2.742 ± 0.081) sin4 b)◦ day−1. The height of CBPs in the SDO/AIA 19.3 nm channel was found to be ≈6500 km. Meridional motion is pre- dominantly poleward for all latitudes, while solar velocity residuals show signs of torsional oscillations. Horizontal Reynolds stress was found to be smaller than in similar works, but still showed transfer of angular momentum towards the solar equator. Conclusions. Most of the results are consistent with Doppler measurements rather than tracer measurements. The fairly small calcu- lated value of horizontal Reynolds stress might be due to the particular phase of the solar cycle. Accuracy of the calculated rotation profile indicates that it is possible to measure changes in the profile as the solar cycle evolves. Analysis of further SDO/AIA CBP data will also provide a better understanding of the temporal behaviour of the rotation velocity residuals, meridional motions, and Reynolds stress. Key words. Sun: rotation – Sun: corona – Sun: activity

1. Introduction measurements. Analyses of tracer data show that meridional flow is going out of the centre of activity (Howard & Gilman Studies of the solar rotation profile, torsional oscillations, and 1986; Wöhl & Brajša 2001), while Doppler measurements usu- meridional velocities are based on the specific features traced ally show poleward meridional flow for all latitudes (Duvall on or above the photosphere or on Doppler measurements. The 1979; Hathaway 1996). Of course, there have been studies that oldest known tracers for measuring the solar rotation profile are show the opposite. For example, Howard (1991) pointed out sunspots, which have been used for a long time (Newton & that solar plages show flow towards the centre of solar activity, Nunn 1951; Howard et al. 1984; Balthasar et al. 1986; Brajša unlike other tracer measurements. In contrast to other Doppler et al. 2002a; Sudar et al. 2014). The biggest advantage of measurements, Pérez Garde et al. (1981) found motion towards sunspots is that they have been observed for more than a cen- the equator from their analysis of Doppler data. Olemskoy & tury. Coronal bright points (CBPs) have also been used very fre- ff Kitchatinov (2005) have pointed out that for tracer measure- quently by using the data obtained by di erent satellites. For ments it is critical to take into account the distribution of the trac- example, Brajša et al. (2001, 2002b, 2004), Vršnak et al. (2003), ers in latitude in order not to detect false flows. Recently, Sudar Wöhl et al. (2010) used SOHO/EIT data, Hara (2009)analysed / et al. (2014) have analysed sunspot group data from Greenwich Yohkoh SXT measurements, while Kariyappa (2008) used both Photoheliographic Results and, by using the arguments from Yohkoh and Hinode data. Recently, Sudar et al. (2015)haveused / Olemskoy & Kitchatinov (2005), find that the meridional flow SDO AIA measurements in the 19.3 nm channel. is towards the centre of solar activity. Doppler measurements give similar results for rotation (Howard & Harvey 1970; Ulrich et al. 1988; Snodgrass & Based on Doppler data, Howard & Labonte (1980) reported Ulrich 1990), but analyses of meridional motions and tor- that the Sun is a torsional oscillator. This was later confirmed sional oscillation differ significantly between tracer and Doppler by Ulrich et al. (1988), again with Doppler measurements, and Howe et al. (2000) with helioseismic measurements. While  Table 1 is only available at the CDS via anonymous ftp to many later papers found the torsional oscillation pattern in such cdsarc.u-strasbg.fr (130.79.128.5)orvia measurements, Sudar et al. (2014) were unable to detect any- http://cdsarc.u-strasbg.fr/viz-bin/qcat?J/A+A/587/A29 thing like it in 150 yr of sunspot group data.

Article published by EDP Sciences A29, page 1 of 6 A&A 587, A29 (2016)

Tracer data is very useful for the analysis of horizontal because of misidentification in subsequent images or some sim- Reynolds stress because both velocity components can be mea- ilar problems. It is quite common to filter out such outliers by sured separately (Schröter 1985). There are a number of papers selecting a fixed range of acceptable rotational velocity (Brajša (Ward 1965; Schröter & Wöhl 1976; Gilman & Howard 1984; et al. 2002b; Vršnak et al. 2003; Sudar et al. 2014, 2015). This Pulkkinen & Tuominen 1998; Vršnak et al. 2003; Sudar et al. approach neglects the fact that the solar rotation varies with lati- 2014) that have found the value of Reynolds stress in agree- tude and that such a fixed cut-off does not have a uniform effect ment with transfer of the angular momentum towards the equator on all latitudes. This in turn can affect the calculated rotation which could explain the observed solar rotation profile. profile. Although, this effect is probably negligible for the solar Sudar et al. (2015)usedSDO/AIA 19.3 nm channel to trace rotation, it might create problems for derived quantities such as CBPs for two days. Their results show that the combination of rotation velocity residuals or Reynolds stress. high cadence/high resolution satellite measurements can provide Brajša et al. (2001)andWöhl et al. (2010) adopted a dif- a wealth of data that can be used to analyse variations of the so- ferent, two-step approach where they first applied the fixed fil- lar rotation profile and all the associated phenomena mentioned ter, calculated the solar rotation profile, and then eliminated all above. Analysis of meridional flow, torsional oscillations, and measurements which differed by more than 2◦ day−1 from the horizontal Reynolds stress with SDO/AIA CBP data is the main calculated profile. Finally, the new profile was calculated with a goal of this paper. truncated dataset. This approach takes into account the variation of the rotation with latitude and performs a cut-off on rotation velocity residuals. 2. Data and reduction methods In an attempt to remove all arbitrariness from the procedure, we also developed a method which removes the outliers based In this work we used measurements from Atmospheric Imaging on the rotation velocity residuals. The method we used is based Assembly (AIA) instrument which is on board the Solar on interquartile range. First we calculate solar rotation profile Dynamics Observatory (SDO) satellite (Lemen et al. 2012). from all data, We used a similar procedure to that used in our previous pa- per (Sudar et al. 2015) to obtain CBP positions. The segmenta- tion algorithm is a modification of similar algorithms described ω(b) = A + B sin2 b + C sin4 b, (1) in McIntosh & Gurman (2005)andMartens et al. (2012). In Table 1, available at CDS, we provide the following informa- where b is the latitude, and then we calculate rotation velocity tion. Column 1 lists the Julian date of each observation, Col. 2 residuals. Then we determine the lower quartile, Q1, and upper contains the identification number of CBP, Cols. 3 and 4 give the quartile, Q , for the rotation velocity residuals distribution. We y 3 x and coordinates for each CBP in pixels, respectively. exclude all datapoints outside of the range In order to obtain better accuracy than in Sudar et al. (2015) where we used observational data from 2 days, in this work we − − , + − , analysed more than 5 months of AIA/SDO observations with a [Q1 k(Q3 Q1) Q3 k(Q3 Q1)] (2) 10-min cadence from 2011 January 1 to 2011 May 19. We re- moved data points near the limb (>0.95 R) in order to avoid po- where we have chosen k = 3.5, which removes the so-called tential problems with inaccuracies in position for those points. hard outliers. With the reduced dataset we calculate solar rota- Choosing only CBPs with ten or more data points to calculate tion profile again and repeat the process iteratively until no dat- velocities by linear fitting, we obtained 82 341 velocity measure- apoints are removed by interquartile criterion. In each iteration ments, which were converted from synodic to sidereal velocities we removed the outliers by looking at meridional velocity dis- (Skokic´ et al. 2014). tribution with the same method. The whole process is finished In Sudar et al. (2015) we showed a change in position over after only a few iterations. time of one CBP in Fig. 2. Apart from the trend line, CBPs also Since CBPs are situated above the photosphere at unknown exhibit apparently random fluctuations around the fitted line. height, we are actually measuring their apparent (projected) he- Such fluctuations might be the result of displacements associated liographic coordinates (Roša et al. 1995, 1998). To correct this with the evolution of CBP photospheric footpoints (Karachik problem we considered that the solar rotation profile is invari- et al. 2014). Another cause might be due to image pixelation or ant to the central meridian distance (CMD). We divided the so- changes in intensity distribution of CBPs. Given that the spatial lar disk into bins that are 10◦ wide in CMD and calculated the / ≈ / resolution of SDO AIA is 0.6 pixel, we can estimate that the rotation profile for each bin. We obtained a series of rotation error in position induced by CBP apparently changing position ffi ◦ profile coe cients – Ai, Bi,andCi – that can be compared with by 1 pixel is about 0.04 in solar coordinates around the equa- theprofileinthe−5◦ to 5◦ CMD range defined by coefficients tor. This is much smaller than observed by Sudar et al. (2015). A, B,andC. We can calculate these coefficients for a num- The error in velocity between two subsequent images would be berofdifferent heights above photosphere and request that the −1 less than 1 m s . In our case, where we actually make a straight function line fit through positions measured in at least ten images where the same CBP is detected, the error is even smaller than that.   π/2   Therefore, we assume that the observed fluctuations in CBP po- 2 4 2 δ = Ai − A + (Bi − B)sin b + (Ci − C)sin b db, sition are most likely caused by the evolution of CBP photo- 0 spheric footpoints described by Karachik et al. (2014). Since i these fluctuations do not have a preferred direction on the so- (3) lar surface, we can assume that this effect averages out with a large number of data points. is lowest for some trial height, h. The integral is taken from the With such a large number of data points obtained by an auto- equator to the pole so that the full profile is taken into account. matic method, it is very likely that some velocities are incorrect This integral can be evaluated since coefficients do not depend

A29, page 2 of 6 D. Sudar et al.: Meridional motions and Reynolds stress from SDO/AIA coronal bright points data

2000 14.5

1800 14.0

1600 13.5

1400 13.0

] 12.5 1200 -1

δ 12.0 day

1000 ° [

ω 11.5 800 11.0 600 10.5 400 10.0 200 9.5 0 2000 4000 6000 8000 10000 12000 0 10 20 30 40 50 60 70 80 h [km] b [°] δ Fig. 1. Calculated as a function of height, h, clearly shows a minimum Fig. 2. Average solar rotation profile in 2◦ bins of latitude, b,isshown value at around 6500 km above the photosphere. with filled black squares and error bars. The best fit solar rotation profile is shown with a solid black line. on the latitude, b, so function δ becomes   π π √ 2 δ = wA (Ai − A) + wA wB (Ai − A)(Bi − B) ffi 2 i 4 i i profile (Eq. (1)) to 80 966 measurements for the coe cients we i obtain A = 14.4060±0.0051◦ day−1, B = −1.662±0.050◦ day−1, π  √  ◦ −1 +3 w − 2 + w w − − and C = −2.742 ± 0.081 day . It is important to point out that Bi (Bi B) Ai Ci (Ai A)(Ci C) 16  previous studies needed decades of measurements to achieve this 5π √ 35π level of accuracy. + w w (B − B)(C − C) + w (C − C)2 , (4) 32 Bi Ci i i 256 Ci i As Snodgrass & Howard (1985) explained, it is not straight- forward to compare the results of the solar rotation profile from where we have introduced weights for coefficients w , w ,and 2 Ai Bi different sources when expressed as an expansion series of sin b w , which are calculated from their errors obtained by fitting Ci (Eq. (1)). To avoid a crosstalk problem between coefficients, it the solar rotation profile in each CMD bin. This height correc- is better to express the result as Gegenbauer polynomials, which tion procedure was performed together with the iterative outlier are orthogonal on the disk. Our solar rotation profile, expressed removal process described above. with Gegenbauer polynomials is given by the coefficients AG = By following the reasoning in Roša et al. (1995) we can 13.8386◦ day−1, B = −0.698◦ day−1,andC = −0.131◦ day−1. transform the apparent coordinates into the deprojected ones by G G Our result for the rotation profile is the most similar to the assuming that CBP are at some height, h. This task is performed ffi in polar coordinates obtained from pixel coordinates (Roša et al. value calculated by Hara (2009) who found the coe cients to be A = 14.39◦ day−1, B = −1.91◦ day−1,andC = −2.45◦ day−1, 1995) so that both heliographic coordinates, CMD and latitude, ffi = . ◦ −1 δ or expressed with Gegenbauer coe cients AG 13 80 day , are corrected for height. We can then simply plot as a func- = − . ◦ −1 = − . ◦ −1 tion of h and from the minimum detect the best fit height. This BG 0 709 day ,andCG 0 117 day . Hara (2009) analysed X-ray bright points observed by the Yohkoh soft X-ray plot is given in Fig. 1. We can clearly see that the minimum − of function δ is located around 6500 km giving us the average telescope in the period 1994 1998. This time period starts close height of CBPs seen in the SDO/AIA 19.3 nm channel. By fit- to the end of cycle 22 and ends soon after the beginning of cy- ting the parabolic function to δ(h) we get the average height, cle 23 (see Table 1 in Brajša et al. 2009). h = 6331 ± 239 km. Perhaps the similarity between our results and those of Hara In the final run after all the filtering and with the best fit (2009) are related to the low solar activity in both works. For height we had 80 966 velocities in our dataset. For the analysis example, Brajša et al. (2004) found a slightly higher value of = . ± . ◦ −1 of rotation velocity residuals and meridional motion we trans- the equatorial rotation A 14 454 0 027 day in the pe- − − formed the velocities to units of m s 1. We calculated meridional riod from 1998 1999 which is closer to the solar activity max- velocities on the southern hemisphere with v = −∂b/∂t and imum. Wöhl et al. (2010) found even faster equatorial rotation mer = . ± . ◦ −1 assigned them symmetrical positive latitudes. This means that a A 14 499 0 006 day with CBP data covering most of negative value of meridional velocity represents motion towards cycle 23 around its activity maximum. If such variations in the the solar equator on both hemispheres. solar rotation profile are indeed due to the changing activity of the sun, given the coefficient uncertainties we calculated above, we should be able to detect and track these changes during the 3. Results solar activity cycle with the expanded SDO/AIA CBP dataset. ω 3.1. Solar rotation profile and rotation velocity residuals In Fig. 2 we show the best fit rotation profile, (b), with a solid black line. We also show average values of ω in 2◦ bins of In our previous paper (Sudar et al. 2015) we estimated that latitude, b, with with black squares and error bars. We see that with 5−6 months of SDO/AIA data we could obtain a sufficient the bin averaged values are fairly well determined up to high number of velocity measurements and that the accuracy of the latitudes (70◦), which is very promising for further CBP studies solar rotation profile would be comparable with the most accu- based on the SDO/AIA data. The small size of the error bars also rate tracer results obtained so far. Fitting the standard rotation illustrates how well the rotation profile is determined.

A29, page 3 of 6 A&A 587, A29 (2016)

4000 25.0

3500 20.0

3000 15.0

2500 ] 10.0 -1

n 2000 5.0 [m s

1500 rot v

Δ 0.0 1000 -5.0 500 -10.0 0 -1000 -800 -600 -400 -200 0 200 400 600 800 1000 -15.0 Δ -1 vrot [m s ] 0 10 20 30 40 50 60 70 80 ° −1 b [ ] Fig. 3. Distribution of rotation velocity residuals, Δvrot,in20ms bins. ◦ Fig. 4. Rotation velocity residuals, Δvrot,in5 bins in latitude, b,are shown with a thick black solid line and filled squares with error bars. Values calculated with bins of the same width, but with the first bin in Rotation velocity residuals, Δvrot, are calculated by subtract- the range 0◦–2.5◦, are shown with a grey solid line and filled circles ing the actual rotation velocity of each CBP from the mean pro- Δv ffi with error bars. rot averages in bins with constant number of data filegivenbythecoe cient of the fit above. The residuals are points (n = 5000) are shown with a dotted line and stars with error bars. further transformed from units of ◦ day−1 to m s−1 where we took into account the latitude of each CBP. In Fig. 3 we show a 4000 distribution of rotation velocity residuals, Δvrot.SinceΔvrot was used to eliminate the outliers, it is important to check whether 3500 there are any unusual features in their distribution, which would indicate that something went wrong with our procedure. The dis- 3000 tribution in Fig. 3 looks fairly normal and well behaved so we 2500 assume that the method used is acceptable.

Tlatov & Pevtsov (2013) proposed an interesting idea that n 2000 the torsional oscillation pattern, associated with rotation veloc- 1500 ity residuals, could – at least partially – be an artefact of binning in latitude b. The authors have been successful in simulating the 1000 torsional oscillation pattern by assuming drifting of the tracers towards the solar equator during the solar activity cycle. This is 500 most notable for sunspots with a characteristic butterfly diagram, 0 but could also be visible for CBPs. Moreover, the authors sug- -1500 -1000 -500 0 500 1000 1500 gest that this effect is present in Doppler and helioseismology -1 vmer [m s ] measurements. v −1 In Fig. 4 we show the values of rotation velocity residuals, Fig. 5. Distribution of meridional velocities, mer,in20ms bins. ◦ Δvrot, grouped into 5 bins of latitude b with a thick black solid line and black squares with error bars. In order to address the though the errors are becoming large in that region of latitudes. problem of binning in latitude, we also calculated Δvrot averages in shifted bins, where the first bin is in the range from 0◦ to 2.5◦ This type of result is more consistent with Doppler measure- and all subsequent bins are 5◦ wide (shown with a grey solid ments than with sunspot data where Sudar et al. (2014) found no line and filled circles with error bars in Fig. 4). In addition, we stable pattern that resembles torsional oscillations. calculated Δvrot averages in bins with a constant number of data = points (n 5000) and show the results in the same graph with a 3.2. Meridional velocities dotted line and stars with error bars. We also calculated average latitude, b¯, instead of using a middle value of b for each bin. In order not to detect false meridional flows due to uneven dis- This process should alleviate the binning problem described by tribution of tracers across different latitudes, it is necessary to Tlatov & Pevtsov (2013), because the value of b¯ is much more assign calculated velocities to the latitude of the first measure- appropriate in the case of uneven distribution of data points in ments of position for each CBP (Olemskoy & Kitchatinov 2005; the bin. Sudar et al. 2014). From Fig. 4 we can conclude that all three of the binning As with Δvrot, it is wise to take a look at the distribu- techniques we used show practically the same behaviour. We tion of meridional velocities, vmer (Fig. 5), because they were also note that the values of b¯ do not differ significantly from the also used in outlier identification and elimination from the raw middle value of b for each bin. The difference is lower than 0.2◦ dataset. Again, there are no unexpected features in the distribu- for all but one bin. This suggests that the distribution of tracers in tion, which leads us to believe that the method used was valid each bin is not far from uniform. In Fig. 4 we see positive values and correctly implemented. ◦ ◦ of Δvrot averages in the range between 10 and 20 , which might In Fig. 6 we show average meridional velocity, vmer,asa correspond to a branch of torsional oscillation pattern towards function of latitude, b. As in the case with Δvrot,weusedthree the equator. Another branch might be visible above 50◦,even different binning techniques and show the results with the same

A29, page 4 of 6 D. Sudar et al.: Meridional motions and Reynolds stress from SDO/AIA coronal bright points data

60.0 3000.0 2500.0 50.0 2000.0 40.0 1500.0

] 1000.0 ] -1

30.0 -2

s 500.0 2 [m s 20.0 0.0 mer q [m v -500.0 10.0 -1000.0 -1500.0 0.0 -2000.0 -10.0 -2500.0 0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80 b [°] b [°] ◦ ◦ Fig. 6. Average meridional velocity, vmer in 5 bins of latitude, b,are Fig. 7. Horizontal Reynolds stress in bins of 10 in latitude, b,isshown shown with a thick black solid line and filled squares with error bars. with filled black squares with error bars. Values calculated with bins of the same width, but with the first bin in the range 0◦–2.5◦, are shown with a grey solid line and filled circles with v error bars. mer averages in bins with constant number of data points show a positive average value, while the horizontal Reynolds (n = 5000) are shown with a dotted line and stars with error bars. stress is negative. If we take into account that the average of the product is not equal to the product of averages, vmerΔvrot  vmer Δvrot , we see that the negative Reynolds stress we ob- symbols as in Fig. 4. From the image we can see that the merid- tained actually means that the two velocities are not independent, ional velocity is almost always positive, meaning towards the but correlated. poles, for all latitudes. This result is in contrast to what was found with sunspot groups (Sudar et al. 2014) where the authors detected flow towards the centre of solar activity for all latitudes 4. Summary and conclusion and for all phases of the solar cycle. Predominantly poleward flow was not found in other works dealing with tracers (Howard By using just under six months of SDO/AIA observations we & Gilman 1986; Wöhl & Brajša 2001; Vršnak et al. 2003). have calculated the solar rotation profile with an accuracy com- On the other hand, poleward flow for all latitudes has been parable to other tracer measurements that needed a much longer detected by using the Doppler method (Duvall 1979; Hathaway time span of observations. The calculated solar rotation profile 1996). There is a small indication that for latitudes near the equa- and comparison with other works indicates that our result is con- tor the flow is towards the equator. Snodgrass & Dailey (1996) nected with a low solar activity in the observed phase of the cur- already reported that this feature is present in their analysis of rent solar cycle. Further work with more SDO/AIA data might Mt. Wilson magnetograms. Moreover, they found that this low provide some valuable insight into the behaviour of the solar ro- latitude behaviour actually changes over the course of the so- tation during the solar cycle. lar cycle. It will be very interesting to check if this behaviour is We found that CBPs observed by SDO/AIA in the 19.3 nm present in the expanded SDO/AIA CBP dataset. channel are located at an average height of ≈6300 km above the solar photosphere. This is slightly lower than in previous stud- ies: Simon & Noyes (1972) ≈11 000 km, Brajša et al. (2004) 3.3. Horizontal Reynolds stress 8000−12 000 km, and Hara (2009) ≈12 000 km. On the other The horizontal Reynolds stress is simply defined as a product hand, Karachik et al. (2006) suggest the value of 80 000 km, xiv λ of rotation velocity residuals and meridional velocities averaged which is the height at which the Fe 195 line forms (Zhang over longitudes for a given latitude band: & et al. 2000). Rotation velocity residuals show indications of torsional oscillations and further studies of the evolution of ob- q = Δvrotvmer . (5) served features might be very helpful for comparison with other methods. In our convention, if the value of q is negative it means that the Meridional velocities are almost always towards the so- angular momentum is transported towards the solar equator. lar poles, which is the direction often seen in helioseismology In Fig. 7 we show the value of the horizontal Reynolds stress, measurements (Zhao & Kosovichev 2004; González Hernández q, as a function of latitude, b, with black squares. The bins are et al. 2008, 2010). Observations of sunspot groups, on the other 10◦ wide and the error bars are also shown. We can see that the hand, show a different meridional velocity pattern (Sudar et al. Reynolds stress is zero for almost all latitudes, b. The only no- 2014). However, Sudar et al. (2014) have pointed out that merid- table exception is the bin at 25◦ where q ≈−1500 m2 s−2 and ional velocity residuals in helioseismology measurements show possibly the bin at 15◦. Sudar et al. (2014) found the minimum a striking similarity with sunspot groups observations. The dif- at the same latitude, but about twice as deep. They also found ference between CBPs and sunspot groups can be explained with negative values of q for all latitudes below 35◦. Latitudes above similar arguments to those in Sudar et al. (2014) who have sug- 35◦ were out of reach for sunspot groups measurements. gested that sunspot observations show motions related to active The minimum at 25◦ seems unexpected when we look at the regions, while the mostly poleward flow is observed outside of behaviour of average meridional velocity, vmer, and rotation ve- those regions (Zhao & Kosovichev 2004,Fig.5).Oursegmen- locity residuals, Δvrot, at the same latitude. The two velocities tation algorithm has difficulties in detecting CBPs over bright

A29, page 5 of 6 A&A 587, A29 (2016) active regions, so the CBPs results are more similar to time- Hara, H. 2009, ApJ, 697, 980 distance heliosiesmology studies than to sunspot measurements. Hathaway, D. H. 1996, ApJ, 460, 1027 Reynolds stress shows a minimum at around 25◦ in latitude Howard, R. F. 1991, Sol. Phys., 135, 327 Howard, R., & Gilman, P. A. 1986, ApJ, 307, 389 similar to the results from Sudar et al. (2014), but with lower Howard, R., & Harvey, J. 1970, Sol. Phys., 12, 23 ◦ 2 −2 2 −2 magnitude (q(25 ) ≈−1500 m s , compared to ≈−3000 m s Howard, R., & Labonte, B. J. 1980, ApJ, 239, L33 in Sudar et al. 2014). We are not sure whether the reason for this Howard, R., Gilman, P. I., & Gilman, P. A. 1984, ApJ, 283, 373 result is the same as for the meridional velocities or whether this Howe, R., Christensen-Dalsgaard, J., Hill, F., et al. 2000, ApJ, 533, L163 Karachik, N., Pevtsov, A. A., & Sattarov, I. 2006, ApJ, 642, 562 is some peculiarity of the phase of the solar cycle or even the Karachik, N. V., Minenko, E., Sattarov, I., Pevtsov, A. A., & Sherdonov, C. T. whole cycle 24. 2014, Astron. Nachr., 335, 1037 Further work on the expanded SDO/AIA dataset and even Kariyappa, R. 2008, A&A, 488, 297 the possible application of the segmentation algorithm to previ- Lemen, J. R., Title, A. M., Akin, D. J., et al. 2012, Sol. Phys., 275, 17 ous satellite measurements (e.g. from SOHO/EIT) will be very Martens, P. C. H., Attrill, G. D. R., Davey, A. R., et al. 2012, Sol. Phys., 275, 79 McIntosh, S. W., & Gurman, J. B. 2005, Sol. Phys., 228, 285 helpful for our understanding of the dynamics on and above the Newton, H. W., & Nunn, M. L. 1951, MNRAS, 111, 413 photosphere. Such research can be considered complementary to Olemskoy, S. V., & Kitchatinov, L. L. 2005, Astron. Lett., 31, 706 helioseismology measurements which probe the behaviour be- Pérez Garde, M., Vázquez, M., Schwan, H., & Wöhl, H. 1981, A&A, 93, 67 low the solar surface. Pulkkinen, P., & Tuominen, I. 1998, A&A, 332, 755 Roša, D., Vršnak, B., & Božic,´ H. 1995, Hvar Observatory Bulletin, 19, 23 Roša, D., Vršnak, B., Božic,´ H., et al. 1998, Sol. Phys., 179, 237 Acknowledgements. This work has been supported in part by the Croatian Schröter, E. H. 1985, Sol. Phys., 100, 141 Science Foundation under the project 6212 “Solar and Stellar Variability”. It Schröter, E. H., & Wöhl, H. 1976, Sol. Phys., 49, 19 has also received funding from the SOLARNET project (312495, 2013−2017), Simon, G. W., & Noyes, R. W. 1972, Sol. Phys., 26, 8 which is an Integrated Infrastructure Initiative (I3) supported by FP7 Capacities Skokic,´ I., Brajša, R., Roša, D., Hržina, D., & Wöhl, H. 2014, Sol. Phys., 289, Programme. S.S. was supported by NASA Grant NNX09AB03G to the 1471 Smithsonian Astrophysical Observatory and contract SP02H1701R from Snodgrass, H. B., & Dailey, S. B. 1996, Sol. Phys., 163, 21 Lockheed-Martin to SAO. Snodgrass, H. B., & Howard, R. 1985, Sol. Phys., 95, 221 Snodgrass, H. B., & Ulrich, R. K. 1990, ApJ, 351, 309 Sudar, D., Skokic,´ I., Ruždjak, D., Brajša, R., & Wöhl, H. 2014, MNRAS, 439, References 2377 Sudar, D., Skokic,´ I., Brajša, R., & Saar, S. H. 2015, A&A, 575, A63 Balthasar, H., Vázquez, M., & Wöhl, H. 1986, A&A, 155, 87 Tlatov, A. G., & Pevtsov, A. A. 2013, in Fifty Years of Seismology of the Sun Brajša, R., Wöhl, H., Vršnak, B., et al. 2001, A&A, 374, 309 and Stars, eds. K. Jain, S. C. Tripathy, F. Hill, J. W. Leibacher, & A. A. Brajša, R., Wöhl, H., Vršnak, B., et al. 2002a, Sol. Phys., 206, 229 Pevtsov, ASP Conf. Ser., 478, 297 Brajša, R., Wöhl, H., Vršnak, B., et al. 2002b, A&A, 392, 329 Ulrich, R. K., Boyden, J. E., Webster, L., Padilla, S. P., & Snodgrass, H. B. 1988, Brajša, R., Wöhl, H., Vršnak, B., et al. 2004, A&A, 414, 707 Sol. Phys., 117, 291 Brajša, R., Wöhl, H., Hanslmeier, A., et al. 2009, A&A, 496, 855 Vršnak, B., Brajša, R., Wöhl, H., et al. 2003, A&A, 404, 1117 Duvall, Jr., T. L. 1979, Sol. Phys., 63, 3 Ward, F. 1965, ApJ, 141, 534 Gilman, P. A., & Howard, R. 1984, Sol. Phys., 93, 171 Wöhl, H., & Brajša, R. 2001, Sol. Phys., 198, 57 González Hernández, I., Kholikov, S., Hill, F., Howe, R., & Komm, R. 2008, Wöhl, H., Brajša, R., Hanslmeier, A., & Gissot, S. F. 2010, A&A, 520, A29 Sol. Phys., 252, 235 Zhang, M., & et al. 2000, in IAU Joint Discussion, 7 González Hernández, I., Howe, R., Komm, R., & Hill, F. 2010, ApJ, 713, L16 Zhao, J., & Kosovichev, A. G. 2004, ApJ, 603, 776

A29, page 6 of 6 4.5 Paper V: Meridional Motion and Reynolds Stress from Debrecen Photoheliographic Data

The following article is published in Solar Physics, Volume 292, Issue 86, pp. 1-13 (2017). It is reproduced by the permission of Springer. The .pdf document is avail- able in the online version of this journal from 19th June 2017, and hereafter. My own contribution to this work was 20%.

The main aims of this paper are presented in Sec. 1. The main conclusions of this paper are listed here. The motion of sunspot groups obtained from the DPD data, as a continuation of the GPR data set, was analysed for the 1974 – 2016 time period. It is shown that the calculated solar rotation profile is in agreement with the results of other authors obtained on GPR data. Therefore, one of the conclusions of this paper is that DPD catalogue could be considered as an adequate successor of the GPR dataset. Another conclusion is connected with the investigation of the meridional motions. The observed meridional motions show the direction towards the zone of solar activity. This means that the latitude at which the flow becomes zero is the centre of solar activ- ity defined by the latitudinal distribution of sunspots for each phase of the solar activity cycle, which corresponds approximately to the absolute latitude of 15◦. Obtained be- haviour of the meridional motions is asymmetrical around the equator, being oriented from equator towards the centre of solar activity, as well as from the poles towards the centre of solar activity. The difference in measured meridional flow in comparison to Doppler measurements and some other tracer measurements is interpreted as a conse- quence of different flow patterns inside and outside of active regions. The different observations of the meridional motion of tracers, towards the centre of activity or oppo- sitely, are suggested to be a consequence of the assumption that it is irrelevant to which latitude one assigns the observed meridional velocity Olemskoy & Kitchatinov (2005). The third most important conclusion of this paper comes from a statistically sig- nificant dependence of meridional motion on rotation velocity residuals. It confirms the transfer of angular momentum towards the equator, as well as calculated horizontal Reynolds stress. The latitudinal behaviour of the horizontal Reynolds stress shows the transport of angular momentum towards the equator everywhere, increasing in absolute value from the equator to a possible maximum value somewhere around 25 – 40◦, which is also consistent with earlier studies based on sunspot data.

74 Solar Phys (2017) 292:86 DOI 10.1007/s11207-017-1105-4

Meridional Motion and Reynolds Stress from Debrecen Photoheliographic Data

Davor Sudar1 · Roman Brajša1 · Ivica Skokic´2 · Ivana Poljanciˇ cBeljan´ 3 · Hubertus Wöhl4

Received: 2 February 2017 / Accepted: 3 May 2017 © Springer Science+Business Media Dordrecht 2017

Abstract The Debrecen Photoheliographic Data catalogue is a continuation of the Green- wich Photoheliographic Results providing daily positions of sunspots and sunspot groups. We analyse the data for sunspot groups focussing on meridional motions and transfer of angular momentum towards the solar equator. Velocities are calculated with a daily shift method including an automatic iterative process of removing the outliers. Apart from the standard differential rotation profile, we find meridional motion directed towards the zone of solar activity. The difference in measured meridional flow in comparison to Doppler mea- surements and some other tracer measurements is interpreted as a consequence of different flow patterns inside and outside of active regions. We also find a statistically significant dependence of meridional motion on rotation velocity residuals confirming the transfer of angular momentum towards the equator. Analysis of horizontal Reynolds stress reveals that the transfer of angular momentum is stronger with increasing latitude up to about 40◦,where there is a possible maximum in absolute value.

Keywords Sunspots · Rotation · Velocity fields · Photosphere

B D. Sudar [email protected] R. Brajša [email protected] I. Skokic´ [email protected] I. Poljanciˇ c´ Beljan [email protected] H. Wöhl [email protected]

1 Hvar Observatory, Faculty of Geodesy, Kaciˇ ceva´ 26, University of Zagreb, 10000 Zagreb, Croatia 2 Astronomical Institute of the Czech Academy of Sciences, Fricovaˇ 298, 25165 Ondrejov,ˇ Czech Republic 3 Department of Physics, University of Rijeka, Radmile Matejciˇ c´ 2, 51000 Rijeka, Croatia 4 Kiepenheuer-Institut für Sonnenphysik, Schöneckstr. 6, 79104 Freiburg, Germany 86 Page 2 of 13 D. Sudar et al.

1. Introduction

The Debrecen Photoheliographic Data (DPD) catalogue was started as a continuation of the Greenwich Photographic Results (GPR). Royal Greenwich Observatory ceased its pho- toheliographic program in 1977 and the International Astronomical Union commissioned Debrecen Observatory, Hungary, to continue the project (Wayman, 1980). The DPD cata- logue is also supplemented by solar images from other observatories (for details see Baranyi, Gyori,˝ and Ludmány, 2016) to fill in the gaps in the Debrecen observations. The GPR dataset represents one of the most valuable resources in studying the behaviour of the Sun over long time periods. It comes as no surprise that various portions of the dataset were used in a large number of papers studying solar rotation (Newton and Nunn, 1951; Ward, 1965, 1966; Balthasar and Wöhl, 1980;Arevaloet al., 1982; Balthasar, Vázquez, and Wöhl, 1986;Brajšaet al., 2002, 2004; Ruždjak et al., 2004, 2005). Often the dataset was ex- tended into the future by using observations by the Solar Observing Optical Network of the United States Air Force/National Oceanic and Atmospheric Administration (SOON/USAF/ NOAA) (Pulkkinen and Tuominen, 1998a; Javaraiah, 2003; Zuccarello and Zappalá, 2003; Javaraiah, Bertello, and Ulrich, 2005;Brajšaet al., 2007; Javaraiah, 2010; Sudar et al., 2014). Pulkkinen and Tuominen (1998a) also extended the analysis into the past by using observations made by Carrington and Spörer. Observations of sunspots from other obser- vatories were also used: e.g., Kanzelhöhe (Lustig, 1983), Mt. Wilson (Gilman and Howard, 1984), Mitaka (Kambry and Nishikawa, 1990), Kodaikanal (Gupta, Sivaraman, and Howard, 1999), and Abastumani (Khutsishvili, Gigolashvili, and Kvernadze, 2002). The transport of angular momentum towards the solar equator, necessary for maintaining the observed solar differential rotation profile, is often attributed to Reynolds stresses which result from mutual dependence between meridional motions and rotation velocity residuals (Rüdiger, 1980; Canuto, Minotti, and Schilling, 1994; Pulkkinen and Tuominen, 1998b). Motions of only a few m s−1 in both velocity components are sufficient to generate hori- zontal Reynolds stress of the order of several 103 m2 s−2, which is sufficient to maintain the observed solar rotation profile (Schröter, 1985, and the references therein). Actual observa- tions show the required value of several 103 m2 s−2 for horizontal Reynolds stress (Ward, 1965;Belvedereet al., 1976; Schröter and Wöhl, 1976; Gilman and Howard, 1984; Pulkki- nen and Tuominen, 1998b;Vršnaket al., 2003; Sudar et al., 2014). Ward (1965), Gilman and Howard (1984), Pulkkinen and Tuominen (1998b), Vršnak et al. (2003), and Sudar et al. (2014) also showed that the absolute value of horizontal Reynolds stress increases with lat- itude up to about 30◦ where there is a possible maximum. Velocity components of the Reynolds stress (meridional motions and rotation velocity residuals) are also a subject of investigation. Howard and Labonte (1980) found a cyclic pattern of alternating faster and slower rotation bands with a period of ≈ 11 years. Further confirmation of these torsional oscillations was reported by Ulrich et al. (1988), Howe et al. (2000), Haber et al. (2002), and Basu and Antia (2003). Recently, Sudar et al. (2014)tried to find a similar pattern in sunspot groups velocity data, but were unable to detect such a signal when folding all solar cycles to one phase diagram. Doppler measurements show meridional flow directed towards solar poles on all latitudes (Duvall, 1979;Howard,1979; Hathaway, 1996; Zhao and Kosovichev, 2004; Kosovichev and Zhao, 2016). Most of the time the flow was of the order of 20 m s−1, but episodes of significantly larger flow have also been detected (Hathaway, 1996). Sometimes almost no meridional flow (Lustig and Wöhl, 1990) or even opposite flow (Pérez Garde et al., 1981) was detected in Doppler data. Various tracers, such as sunspots, sunspot groups, small magnetic features, coronal bright points (CBPs) and solar plages, were also used to measure the meridional motion. Sunspots Meridional Motion and Reynolds Stress Page 3 of 13 86 and sunspot groups were most frequently used and most of the results show flow outward from the centre of the solar activity (Tuominen and Virtanen, 1984; Howard and Gilman, 1986;Howard,1991a;Kambryet al., 1991; Wöhl and Brajša, 2001). Recently, however, Sudar et al. (2014) reported flow which is almost exactly the opposite. Similar flow to Sudar et al. (2014) was found by Howard (1991b) by tracing solar plages. Komm, Howard, and Harvey (1993) and Snodgrass and Dailey (1996) used small magnetic features as tracers and found similar, but still in some aspects mutually different, results. Komm, Howard, and Harvey (1993) found poleward flow at all latitudes with a maximum amplitude of about 10 m s−1, while Snodgrass and Dailey (1996) obtained a flow which is directed out of the centre of solar activity with poleward flow at higher latitudes with similar amplitude to that of Komm, Howard, and Harvey (1993). Analysing the motion of CBPs, Sudar et al. (2016) detected poleward meridional flow for all latitudes with an amplitude of about 30 m s−1. The main focus of this paper is to analyse the meridional motion revealed by tracing sunspot groups over the solar surface with DPD, which has never been used before for this purpose. We also attempt to explain the wide variety of results for meridional flow in Section 4 in a consistent way. Another very important aspect of this work is to confirm that Reynolds stress is the main generator of the observed solar rotation profile as already suggested by a number of previous theoretical and empirical papers.

2. Data and Reduction Methods

Our dataset is created from positions and times of sunspot groups measured in DPD1 (Baranyi, Gyori,˝ and Ludmány, 2016;Gyori,˝ Ludmány, and Baranyi, 2017) during a pe- riod from 1974 to 2016. Positions of sunspot groups are typically determined once per day. We use the daily change in position of each sunspot group to calculate both meridional and rotational speeds: CMD ω = , (1) rot t b ω = , (2) mer t where CMD is the difference between central meridional distances (CMD) and b is the difference in latitude of two consecutive positions of the sunspot group in time t (usually 1day). This results in 59 090 data points for each speed. Rotation speeds are transformed from synodic to sidereal values (Skokic´ et al., 2014). Assigning particular latitude, b, to each ve- locity is not as straightforward as it seems. With two measurements of position to calculate one velocity, we could use the latitude of the first or the second measurement. Even aver- age latitude seems like a possible candidate. However, as Olemskoy and Kitchatinov (2005) pointed out, an uneven distribution of sunspots (or any other kind of tracer) in latitude cre- ates a problem with some of the possible choices. They have shown that the gradient of the latitudinal distribution can create false meridional flows if the latitude of the second mea- surement is used. The simplest way to avoid this problem is to use the latitude of the first measurement and this is the choice we make (Sudar et al., 2014, 2015, 2016). Determination of the position of a sunspot on the limb can introduce large errors due to projections effects. In addition, Arevalo et al. (1982) found that the calculated rotation

1See http://fenyi.solarobs.csfk.mta.hu/en/databases/DPD/. 86 Page 4 of 13 D. Sudar et al.

Figure 1 Best fit solar differential rotation profile, ω(b), is shown with the solid line. Solid squares with error bars are bin-averaged values of the rotation velocity where bins are ◦ 2 wide in latitude.

profile changes, depending on the cut-off longitude. Wöhl (1983) also reported that the de- termined rotation velocities are larger for smaller cut-off longitudes. Such changes were at- tributed to Wilson depression by Balthasar and Wöhl (1983). Therefore we limit the dataset to ± 58◦ in CMD which alleviates these problems (Stark and Wöhl, 1981; Balthasar et al., 1986). Nevertheless, outliers resulting from misidentification of sunspot groups in subse- quent measurements or other errors are still present, so we use an iterative filtering method similar to the one used in Sudar et al. (2016). In the first step we calculate the solar rotation profile: ω(b) = A + B sin2(b), (3) where b is the latitude, and then we calculate the rotation velocity residuals by subtracting the individual rotational speeds from the average rotation profile given by Equation (3). In the next step we calculate the lower quartile, Q1, and upper quartile, Q3, for rotation velocity residuals and finally remove the so-called hard outliers which lie outside of the range:   Q1 − k(Q3 − Q1), Q3 + k(Q3 − Q1) , (4) where we choose k = 3.5. Since outliers could potentially influence the calculated rotation profile, we use the remaining dataset to recalculate the solar rotation profile. The new rota- tion profile changes the values of rotation velocity residuals which then need to be checked for outliers. This process continues iteratively until no data points are outside of the in- terquartile range (Equation (4)). In the end we obtain a dataset consistent with the calculated rotation profile and with outliers removed. In each iteration we also remove the outliers in meridional velocity using the same form of the interquartile range. The process converges very fast and only four iterations were necessary in our case. Outlier limits turn out to be ± 4.5◦ day−1 and ± 2.1◦ day−1 for rotation velocity residuals and meridional velocities, re- spectively. After removing the outliers we end up with 53 283 data points. The best fit rota- tion profile is shown in Figure 1 where the coefficients of the profile (Equation (3)) are A = 14.5011 ± 0.0081◦ day−1 and B =−2.540 ± 0.073◦ day−1. In the same figure we also show average values of ω(b) in 2◦-wide bins in latitude. The error bars shown in Fig- ure 1 become fairly large for b>35◦, which is a consequence of the fact that sunspots rarely appear above this latitude. In terms of expansion in Gegenbauer polynomials (Snod- ◦ −1 grass and Howard, 1985) the rotation profile coefficients become AG = 13.993 day and ◦ −1 BG =−0.51 day . Meridional Motion and Reynolds Stress Page 5 of 13 86

Figure 2 Meridional flow, vmer as a function of latitude, b, is shown with black squares with error bars. ◦ Averaging bins are 5 wide in latitude. On the left hand side we show all data folded into one hemisphere where meridional velocity is asymmetrically transformed for southern latitudes. In this part of the plot positive values indicate motion towards the poles (+P). On the right hand side of the plot meridional velocities are not transformed and we show both hemispheres. Positive values indicate motion towards the northern solar pole (+N).

In the rest of the paper we use only residual rotation velocities and meridional velocities which are transformed from angular values to their linear counterparts taking into account 8 the latitude of the first measurement. The conversion factor, with R = 6.96 × 10 m, is 140.6ms−1 day(◦)−1, while rotation velocity residuals are additionally multiplied with the cosine of latitude. Additionally, meridional speeds are symmetrised, so that negative value of meridional speed reflects motion towards the equator for both solar hemispheres. This is achieved by transforming calculated meridional speeds with vmer =−∂b/∂t for the southern solar hemisphere, where we assign negative values to b for southern latitudes.

3. Results

In Figure 2 we show the dependence of the meridional flow on latitude, b. Black squares with error bars depict average meridional flow in 5◦ bins of latitude. We can see that for −15◦ < b<15◦ meridional flow is towards the solar poles. In the right hand side of the plot we see that at the equator meridional flow is ≈ 0ms−1 and that the flow is asymmetrical around the equator. At mid-latitudes meridional flow becomes zero again on both hemispheres and probably turns to flow towards the equator at even higher latitudes. Similar behaviour was found by Sudar et al. (2014) who also concluded that the latitude at which the flow becomes zero is the centre of solar activity defined by the latitudinal distribution of sunspots for each phase of the solar activity cycle. On the other hand, by using Doppler line shifts, most researchers found poleward flow of about 20 m s−1 for all latitudes (Duvall, 1979;Howard,1979;Hathaway,1996). Similar values were found by using high-resolution magnetograms (Komm, Howard, and Harvey, 1993) and by applying time–distance helioseismology (Zhao and Kosovichev, 2004). More- over, by using coronal bright points (CBP) as tracers, Sudar et al. (2016) found poleward meridional flows everywhere except at the equator where the flow was zero. This discrep- ancy in meridional flow between sunspot measurements and other observations is discussed in the next section. 86 Page 6 of 13 D. Sudar et al.

Figure 3 Rotation velocity residuals, vrot, as a function of latitude, b, are shown with black squares with error bars. Width of ◦ averaging bins in latitude is 5 . Positive values show faster than average rotation.

Figure 4 Individual observations are shown with dots in the vmer – vrot parameter space. We also show the best linear fit function (Equation (5)) with a solid line.

◦ Average values of the rotation velocity residuals, vrot,inbinsof5 in latitude, b,are shown in Figure 3 with solid squares with error bars. No significant deviation from vrot = 0 can be detected. One important feature of the solar rotation profile is that lower latitudes rotate faster than higher latitudes which implies that most of the angular momentum of the Sun is located in the zones around the equator. Therefore, it is reasonable to assume that there must be some mechanism which transfers angular momentum towards lower latitudes. By studying the relationship between meridional velocities, vmer, and rotation velocity residuals, vrot, we can directly observe if this is what is really happening. The relationship between the two velocities is shown in Figure 4. We also show the best linear fit to the data with the solid line, obtaining the following relation:

−1 vmer = (−0.0876 ± 0.0021)vrot + (1.01 ± 0.34) ms . (5)

The slope in Equation (5) is statistically significant when compared with its uncertainty (relative error ≈2.4%) which shows that on average the two values are not independent. The fact that the slope is negative shows that on average the angular momentum is indeed transferred from higher to lower latitudes. In a previous study, using the sunspot group data from the GPR and the SOON/USAF/NOAA obtained in the period 1878 – 2011, Sudar et al. (2014) found a very similar value for the slope (−0.080 ± 0.002). Reynolds stress is thought to be the main generator of the differential rotation on the Sun (Rüdiger, 1980; Pulkkinen and Tuominen, 1998b) which works against the diffusive decay. Meridional Motion and Reynolds Stress Page 7 of 13 86

Figure 5 Reynolds stress, q =vmervrot,isshownasa function of latitude with black squares with error bars. ◦ Averaging bins are 10 wide in latitude. We also show the best fit function of the form given in Equation (6) with best fit coefficients from Table 1. Shaded area depicts the range of possible values defined by the errors of the best fit coefficients (Table 1).

Table 1 Table of the best fit coefficients (Equation (6)). Coef. Value Rel. error

2 −2 ◦ −1 c1 [m s ( ) ]−166 ± 10 6.2% ◦ −2 c3 [( ) ] 0.00028 ± 0.00011 39.9%

The covariance of meridional velocities and rotation velocity residuals, q =vrotvmer,is the horizontal component of the Reynolds stress tensor and in Figure 5 we show its depen- dence on latitude. Values of q in 10◦ bins in latitude, b, are shown as solid squares with error bars. We can see that the average value of q is negative for all latitudes meaning that there is a net angular momentum transfer towards lower latitudes. This picture might seem to be in conflict with Figures 2 and 3, which show the average behaviour of vmer and vrot with respect to latitude, b. However, we must remember that the average of the product is not equal to the product of the average, i.e. q =vmervrot=vmervrot. Negative values of q actually imply that vmer and vrot are not mutually independent variables and that the relationship between them must be similar to the one given in Equation (5). Sudar et al. (2014) introduced an empirical exponential cut-off function which describes the decreasing trend of q(b) from the equator to higher latitudes and allows for a possi- ble minimum. In this work we simplify the form of this relationship by requiring that the function is perfectly asymmetric around zero (q(b) =−q(−b)) leading to the expression:

2 −c3b q = c1be . (6)

This choice is justified by the appearance of the function q in Figure 10 in Sudar et al. (2014), Figure 15 in Canuto, Minotti, and Schilling (1994), and also by the fact that Sudar −2 et al. (2014) found that the value of parameter e2 in their model was 69 ± 80 m s .The function is plotted in Figure 5 with a thick solid line. In the same plot we also shade the area defined by errors of the coefficients c1 and c3 (Table 1). Using the errors of the coefficients we can also calculate the depth and location of the minimum in the q(b) plot with their respective errors by using the method of error propagation. The minimum is thus located at ◦ 2 −2 bmin = (42.3 ± 8.3) with a value of q(bmin) = (4250 ± 880) m s . A similar result was obtained by Sudar et al. (2014) who suggested that the q(b) relationship has a minimum (q ≈−3000 m s−2) around 25 – 30◦. We do not consider the difference between the two results to be significant, especially considering the scarcity of sunspots at higher latitudes where the minimum is supposed to be. 86 Page 8 of 13 D. Sudar et al.

4. Discussion

The rotation profile calculated in this work is almost identical to the result obtained by Sudar et al. (2014) who also analysed the rotation by tracing sunspot groups albeit for a different time period and different observing stations. So, the DPD series proves to be a rather good continuation of the GPR dataset. For a more thorough comparison of solar rotation profiles obtained by different methods we refer the reader to Wöhl et al. (2010) and Sudar et al. (2015). In Figure 3 we show the average of rotation velocity residuals, vrot, as a function of latitude, b. This plot should not be confused with the torsional oscillation pattern, because torsional oscillations show deviations from the mean velocity in time. When we average these variations over a long time, as we do in Figure 3, such a plot is more indicative of the quality of the fit of the solar rotation profile function (Equation (3)). In this context, we can ◦ say that, apart from the bin at b = 2.5 , the average value of vrot is zero. We are not sure if the slight discrepancy from this rule at b = 2.5◦ is of any significance. The most interesting are the results for the meridional flow (Figure 2) because they are the most controversial. A quote by Hathaway (1996) is perhaps the most illustrative of the problem: “Unfortunately, previous measurements of the Sun’s meridional circulation have produced a bewildering array of results that appear to be of no help at all in constraining theory.” On one hand we have sunspot measurements (and other features closely associated with sunspots such as plages) which show opposite flows on opposite sides of the centre of the solar activity and on the other hand we have Doppler measurements which predomi- nately show poleward motion for all latitudes, regardless of the centre of the solar activity. There are also tracers, such as CBPs (Sudar et al., 2016), which also show average pole- ward motion everywhere of approximately the same amplitude as Doppler measurements. Small magnetic features analysed with different methods by Komm, Howard, and Harvey (1993) and Snodgrass and Dailey (1996) showed different behaviour around the centre of solar activity for meridional motion. So, if we measure the same phenomenon with different techniques, how can we consis- tently interpret these differing results? First of all there is an internal problem with tracer measurements where some authors (Tuominen and Virtanen, 1984; Howard and Gilman, 1986;Howard,1991a;Kambryet al., 1991; Wöhl and Brajša, 2001) reported a flow out of the centre of activity, while others detected a flow towards the centre of activity (Howard, 1991b; Sudar et al., 2014). This difference can probably be explained by an easy-to-make error: assuming that it is irrelevant to which latitude one assigns the observed meridional velocity. Olemskoy and Kitchatinov (2005) demonstrated and Sudar et al. (2014) later ver- ified that it is necessary to assign the latitude of the first measurements due to the uneven distribution of tracers in latitude. Otherwise one would obtain almost exactly opposite and false flow. Technically it is possible to use the latitude of the last measurement of position, but the velocity would have to be weighted taking into account the frequency of tracers at the starting latitude. Since the tracer distribution can also vary during the solar cycle, the second method turns out to be very complicated compared to the neat trick of assigning the latitude of the first measurement. As a consequence we take the results by Sudar et al. (2014) as a reference model for sunspot measurements which are also confirmed by the results for meridional flow in this paper (Figure 2). However, we must point out that Sivaraman et al. (2010), who also used first latitude as a reference point, found a flow which is directed to- wards the solar equator for all latitudes. What we find intriguing in their results is that it appears that meridional flow is not zero at the equator, which is especially noticeable for the Mt. Wilson data they used. On the right hand side of our Figure 2 we can see that the Meridional Motion and Reynolds Stress Page 9 of 13 86 average meridional motion is ≈ 0 which makes the motion clearly asymmetrical on the two solar hemispheres. Small magnetic features were also used as tracers by past authors and although their latitudinal distribution is not the same as the one for sunspots we encounter the same prob- lem. Their distribution overall falls from the equator towards the poles, but it also has peaks around centres of solar activity (Harvey, 1993). Such a distribution gradient can create false flows out of the centre of activity in a similar manner as with sunspots, if proper care of assigning latitude was not taken. In this context we can explain the results by Snodgrass and Dailey (1996), but the meridional flow measured by Komm, Howard, and Harvey (1993) still requires some explanation. First we note that the amplitude of meridional motion in their work was about 10 m s−1 which is substantially lower than most Doppler measure- ments and analysis of CBP data, so it is possible that their result was also influenced just by the overall drop in distribution of tracers towards the poles where the velocity resolution was not sufficient to be influenced by the peaks near centres of activity. Of course, it is also possible that they measured the real flow and that the fairly low amplitude they obtained is consistent with the wide range of possible variations mentioned by Hathaway (1996). Analysing the flow in the upper convective zone with time–distance helioseismology Zhao and Kosovichev (2004) found mostly poleward meridional flow (their Figure 3a). However, when they subtracted the flow from Carrington Rotation 1911 which occurred during the solar minimum year 1996 they obtained the residual meridional flow (their Fig- ure 3b) which in all important aspects looks the same as the meridional flow we see in this work (Figure 2) and in Figure 3 in Sudar et al. (2014). Applying ring-diagram analy- sis to GONG Dopplergrams González Hernández et al. (2008) also found residual merid- ional flows very similar to our meridional flow (their Figures 4 and 5). Moreover, map plots of residual meridional motion in González Hernández et al. (2008)(theirFigure6)and González Hernández et al. (2010) (their Figure 3) look a lot like a similar map plot for meridional motion of sunspot group data in Sudar et al. (2014)(theirFigure2). The key in understanding why residual meridional flow in Doppler measurements looks like total meridional motion in sunspot groups measurements is in appreciating that sunspot groups do not cover all of the solar surface, but are limited to active regions. When Zhao and Kosovichev (2004) subtracted the meridional flow measured during the solar cycle minimum when active regions rarely appear they obtained the residual flow dominated by the motion in and around active regions and that is the area where sunspot groups are found. Therefore, it is perfectly understandable why Zhao and Kosovichev (2004) found that their residual meridional flows in the period from 1997 – 2002 converge towards activity belts with the magnitude of 2 – 8 m s−1 in both solar hemispheres just as we see in our total meridional flow (Figure 2). By analysing sunspot group data we get no information as regards the dominant poleward meridional flow outside of the active regions. Even the analysis of CBP data (Sudar et al., 2016) supports this hypothesis, because detection of CBPs with the segmentation algorithm above bright active regions is almost impossible. As a result, Sudar et al. (2016) found the dominant poleward meridional motion originating outside of the active regions. An alternative to our hypothesis is that the anchor depth of magnetic features plays a role in the meridional flow pattern. Anchor depth has been used as an explanation for differences in rotation rates measured by different tracers and different observing methods. A useful overview of this problem is given in Beck (2000). The rotation rate of surface features are usually connected to depth corresponding to the rotation rate obtained with helioseismology. In general larger features are associated with larger depths, but there are exceptions like for example recurrent sunspot groups. It is logical to assume that the meridional flow of surface features also depicts the motion of deeper layers which is what Sivaraman et al. (2010) 86 Page 10 of 13 D. Sudar et al. concluded based on observing different size sunspot groups. Further work on meridional flow at different depths, possibly in connection with the phase of the solar cycle, is still necessary to shed more light on the subject. Another interesting question is how does this difference between sunspot group data and CBP data affect the observed horizontal Reynolds stress. Sunspot group measurements (Ward, 1965; Gilman and Howard, 1984; Sudar et al., 2014) show a fairly large negative value of q ≈−3000 m2 s−2 located near 30◦ in latitude. A slightly larger value at higher latitude is also found in this work. Canuto, Minotti, and Schilling (1994) already confirmed that this result is in agreement with the theoretical curve using only standard values for parameters in solar conditions. On the other hand, analysis of CBP data by Sudar et al. (2016) yields a highly uncertain value of only q ≈−1500 m2 s−2 around 25◦. The time period covered in Sudar et al. (2016) is less than six months in the rising phase of the solar cycle, but we do not think that the phase of the solar cycle is the cause of such a low value of q in CBP data. Sudar et al. (2014)showedamapofq versus latitude and phase of the solar cycle and found that the locations and depth of the horizontal Reynolds stress, q, are persistent at least from the beginning of the solar cycle until well past the solar activity maximum. In the declining phase of the solar cycle, the latitudinal distribution of sunspots prevents making a firm conclusion about the behaviour of q. So, we suggest that the horizontal Reynolds stress is stronger in and around active regions which means that most of the angular momentum transfer towards the equator takes place in these areas. Another possibility is that the height of the tracers also plays a role, or in other words that Reynolds stress is less pronounced in the layers above the photosphere. This would also imply that CBPs are not firmly rooted to the photosphere and we do not consider this to be very likely.

5. Conclusion

We analyse the motion of sunspot groups obtained by Debrecen Observatory in the period 1974 – 2016. The calculated solar rotation profile is in agreement with other authors es- pecially those investigating also sunspot groups. This shows that the DPD catalogue is an adequate successor of the GPR dataset. The observed meridional motion is typical for the sunspot groups, but different from Doppler and CBP measurements. We suggest that the difference lies in the fact that sunspot groups are located within active regions where the meridional motion is different than out- side of those areas. This idea is supported by residual meridional flow obtained by sub- tracting the dominant poleward meridional flow in Doppler data which then looks almost the same as the total meridional flow we see when analysing sunspot group data. Canuto, Minotti, and Schilling (1994) derived a model of Reynolds stress which is driven by buoy- ancy which acts as a source of convective turbulence. If we take into account that in strong magnetic fields within sunspots convection is inhibited, it might be possible that the velocity components of the horizontal Reynolds stress are also affected and, therefore, show different behaviour in and outside of active regions. The horizontal Reynolds stress calculated in this work confirms the results in all impor- tant aspects obtained by Sudar et al. (2014) who used data from different observatories. The latitudinal behaviour of the horizontal Reynolds stress shows the transport of angular mo- mentum towards the equator everywhere, increasing in absolute value from the equator to a possible maximum value somewhere around 25 – 40◦, which is also consistent with earlier studies based on sunspot data (Ward, 1965;Belvedereet al., 1976; Schröter and Wöhl, 1976; Gilman and Howard, 1984; Pulkkinen and Tuominen, 1998b). By comparing the horizontal Meridional Motion and Reynolds Stress Page 11 of 13 86

Reynolds stress obtained with sunspot group data and CBP data we suggest that larger val- ues of horizontal Reynolds stress, and consequently angular momentum transfer, are found within active regions. Analysis of meridional motion and horizontal Reynolds stress values with CBP data as a function of proximity to active regions might prove or disprove this hypothesis. Measurements of the latitudinal dependence of the covariance of rotation velocity resid- uals and meridional flow presented in this work strongly suggest that the Reynolds stress is the dominant mechanism which explains the observed characteristics of the solar differential rotation. The observed horizontal Reynolds stress is consistent with the transport of angular momentum towards the solar equator both qualitatively (correct sign) and quantitatively (ab- solute value). On the other hand, the model of axisymmetric meridional circulation, which could also produce transfer of the angular momentum towards the equator, is strongly dis- favored because it requires surface meridional motions in the direction towards the solar equator (Gilman, 1981; Stix, 1989; Foukal, 2013), which contradicts the observations in the present and other papers.

Acknowledgements The authors wish to thank the staff of Debrecen Observatory for maintaining and organising the DPD catalogue. This work has been supported in part by the Croatian Science Foundation under the project 6212 “Solar and Stellar Variability”. It has also received funding from the SOLARNET project (312495, 2013 – 2017) which is an Integrated Infrastructure Initiative (I3) supported by FP7 Capacities Programme.

Disclosure of Potential Conflicts of Interest The authors declare that they have no conflicts of interest.

References

Arevalo, M.J., Gomez, R., Vazquez, M., Balthasar, H., Wöhl, H.: 1982, Differential rotation and meridional motions of sunspots from 1874 to 1902. Astron. Astrophys. 111, 266. ADS. Balthasar, H., Vázquez, M., Wöhl, H.: 1986, Differential rotation of sunspot groups in the period from 1874 through 1976 and changes of the rotation velocity within the solar cycle. Astron. Astrophys. 155, 87. ADS. Balthasar, H., Wöhl, H.: 1980, Differential rotation and meridional motions of sunspots in the years 1940 – 1968. Astron. Astrophys. 92, 111. ADS. Balthasar, H., Wöhl, H.: 1983, On the determination of heliographic positions and rotation velocities of sunspots. II – Systematic effects caused by the Wilson depression. Solar Phys. 88, 71. DOI. ADS. Balthasar, H., Lustig, G., Wöhl, H., Stark, D.: 1986, The solar rotation elements i and omega derived from sunspot groups. Astron. Astrophys. 160, 277. ADS. Baranyi, T., Gyori,˝ L., Ludmány, A.: 2016, On-line tools for solar data compiled at the Debrecen observatory and their extensions with the Greenwich sunspot data. Solar Phys. DOI. ADS. Basu, S., Antia, H.M.: 2003, Changes in solar dynamics from 1995 to 2002. Astrophys. J. 585, 553. DOI. ADS. Beck, J.G.: 2000, A comparison of differential rotation measurements – (Invited Review). Solar Phys. 191, 47. DOI. ADS. Belvedere, G., Godoli, G., Motta, S., Paterno, L., Zappala, R.A.: 1976, Surface Reynolds stresses determined from the analysis of facular motions and the maintenance of the sun’s differential rotation. Solar Phys. 46, 23. DOI. ADS. Brajša, R., Wöhl, H., Vršnak, B., Ruždjak, D., Sudar, D., Roša, D., Hržina, D.: 2002, Differential rotation of stable recurrent sunspot groups. Solar Phys. 206, 229. DOI. ADS. Brajša, R., Wöhl, H., Ruždjak, D., Schawinski-Guiton, K.: 2004, Variation of the solar rotation during the activity cycle applying the residual method to Greenwich data. Hvar Obs. Bull. 28, 55. ADS. Brajša, R., Wöhl, H., Ruždjak, D., Vršnak, B., Verbanac, G., Svalgaard, L., Hochedez, J.-F.: 2007, On the solar rotation and activity. Astron. Nachr. 328, 1013. DOI. ADS. Canuto, V.M., Minotti, F.O., Schilling, O.: 1994, Differential rotation and turbulent convection: a new Reynolds stress model and comparison with solar data. Astrophys. J. 425, 303. DOI. ADS. Duvall, T.L. Jr.: 1979, Large-scale solar velocity fields. Solar Phys. 63,3.DOI. ADS. 86 Page 12 of 13 D. Sudar et al.

Foukal, P.V.: 2013, Solar Astrophysics, 3rd, 216. Revised edition. Gilman, P.A.: 1981, Global circulation and the solar dynamo. NASA Spec. Publ. 450. ADS. Gilman, P.A., Howard, R.: 1984, On the correlation of longitudinal and latitudinal motions of sunspots. Solar Phys. 93, 171. DOI. ADS. González Hernández, I., Kholikov, S., Hill, F., Howe, R., Komm, R.: 2008, Subsurface meridional circulation in the active belts. Solar Phys. 252, 235. DOI. ADS. González Hernández, I., Howe, R., Komm, R., Hill, F.: 2010, Meridional circulation during the extended solar minimum: another component of the torsional oscillation? Astrophys. J. Lett. 713, L16. DOI. ADS. Gupta, S.S., Sivaraman, K.R., Howard, R.F.: 1999, Measurement of Kodaikanal white-light images – III. Rotation rates and activity cycle variations. Solar Phys. 188, 225. DOI. ADS. Gyori,˝ L., Ludmány, A., Baranyi, T.: 2017, Comparative analysis of Debrecen sunspot catalogues. Mon. Not. Roy. Astron. Soc. 465, 1259. DOI. ADS. Haber, D.A., Hindman, B.W., Toomre, J., Bogart, R.S., Larsen, R.M., Hill, F.: 2002, Evolving submerged meridional circulation cells within the upper convection zone revealed by ring-diagram analysis. Astro- phys. J. 570, 855. DOI. ADS. Harvey, K.L.: 1993, Properties of emerging bipolar active regions. In: Zirin, H., Ai, G., Wang, H. (eds.) IAU Colloq. 141: The Magnetic and Velocity Fields of Solar Active Regions, Astronomical Society of the Pacific Conference Series 46, 488. ADS. Hathaway, D.H.: 1996, Doppler measurements of the Sun’s meridional flow. Astrophys. J. 460, 1027. DOI. ADS. Howard, R.: 1979, Evidence for large-scale velocity features on the sun. Astrophys. J. Lett. 228, L45. DOI. ADS. Howard, R.F.: 1991a, Cycle latitude effects for sunspot groups. Solar Phys. 135, 327. DOI. ADS. Howard, R.F.: 1991b, The magnetic fields of active regions – part four. Solar Phys. 131, 259. DOI. ADS. Howard, R., Gilman, P.A.: 1986, Meridional motions of sunspots and sunspot groups. Astrophys. J. 307, 389. DOI. ADS. Howard, R., Labonte, B.J.: 1980, The sun is observed to be a torsional oscillator with a period of 11 years. Astrophys. J. Lett. 239, L33. DOI. ADS. Howe, R., Christensen-Dalsgaard, J., Hill, F., Komm, R.W., Larsen, R.M., Schou, J., Thompson, M.J., Toomre, J.: 2000, Deeply penetrating banded zonal flows in the solar convection zone. Astrophys. J. Lett. 533, L163. DOI. ADS. Javaraiah, J.: 2003, Long-term variations in the solar differential rotation. Solar Phys. 212, 23. DOI. ADS. Javaraiah, J.: 2010, Long-term variations in the mean meridional motion of the sunspot groups. Astron. As- trophys. 509, A30. DOI. ADS. Javaraiah, J., Bertello, L., Ulrich, R.K.: 2005, Long-term variations in solar differential rotation and sunspot activity. Solar Phys. 232, 25. DOI. ADS. Kambry, M.A., Nishikawa, J.: 1990, Solar differential rotation derived from sunspot observations. Solar Phys. 126, 89. DOI. ADS. Kambry, M.A., Nishikawa, J., Sakurai, T., Ichimoto, K., Hiei, E.: 1991, Solar meridional motions derived from sunspot observations. Solar Phys. 132, 41. DOI. ADS. Khutsishvili, E.V., Gigolashvili, M.S., Kvernadze, T.M.: 2002, Differential rotation of the sun determined tracing sunspots and oscillations of sunspot tilt angle. Solar Phys. 206, 219. DOI. ADS. Komm, R.W., Howard, R.F., Harvey, J.W.: 1993, Rotation rates of small magnetic features from two- and one-dimensional cross-correlation analyses. Solar Phys. 145,1.DOI. ADS. Kosovichev, A.G., Zhao, J.: 2016, Reconstruction of solar subsurfaces by local helioseismology. In: Rozelot, J.-P., Neiner, C. (eds.) Lecture Notes in Physics 914, Springer, Berlin, 25. DOI. ADS. Lustig, G.: 1983, Solar rotation 1947 – 1981 – determined from sunspot data. Astron. Astrophys. 125, 355. ADS. Lustig, G., Wöhl, H.: 1990, Solar meridional plasma motions from 1982 until 1986. Astron. Astrophys. 229, 224. ADS. Newton, H.W., Nunn, M.L.: 1951, The Sun’s rotation derived from sunspots 1934 – 1944 and additional results. Mon. Not. Roy. Astron. Soc. 111, 413. ADS. Olemskoy, S.V., Kitchatinov, L.L.: 2005, On the determination of meridional flow on the Sun by the method of tracers. Astron. Lett. 31, 706. DOI. ADS. Pérez Garde, M., Vázquez, M., Schwan, H., Wöhl, H.: 1981, Large-scale solar motions as determined by Doppler shift measurements using a linear photodiode array. Astron. Astrophys. 93, 67. ADS. Pulkkinen, P., Tuominen, I.: 1998a, Velocity structures from sunspot statistics in cycles 10 to 22. I. Rotational velocity. Astron. Astrophys. 332, 748. ADS. Pulkkinen, P., Tuominen, I.: 1998b, Velocity structures from sunspot statistics in cycles 10 to 22. II. Latitudi- nal velocity and correlation functions. Astron. Astrophys. 332, 755. ADS. Meridional Motion and Reynolds Stress Page 13 of 13 86

Rüdiger, G.: 1980, Reynolds stresses and differential rotation. I – On recent calculations of zonal fluxes in slowly rotating stars. Geophys. Astrophys. Fluid Dyn. 16, 239. DOI. ADS. Ruždjak, D., Ruždjak, V., Brajša, R., Wöhl, H.: 2004, Deceleration of the rotational velocities of sunspot groups during their evolution. Solar Phys. 221, 225. DOI. ADS. Ruždjak, D., Brajša, R., Sudar, D., Wöhl, H.: 2005, The influence of the evolution of sunspot groups on the determination of the solar velocity field. Solar Phys. 229, 35. DOI. ADS. Schröter, E.H.: 1985, The solar differential rotation – present status of observations. Solar Phys. 100, 141. DOI. ADS. + Schröter, E.H., Wöhl, H.: 1976, Differential rotation and giant cell circulation of solar Ca -network. Solar Phys. 49, 19. DOI. ADS. Sivaraman, K.R., Sivaraman, H., Gupta, S.S., Howard, R.F.: 2010, Return meridional flow in the convection zone from latitudinal motions of umbrae of sunspot groups. Solar Phys. 266, 247. DOI. ADS. Skokic,´ I., Brajša, R., Roša, D., Hržina, D., Wöhl, H.: 2014, Validity of the relations between the synodic and sidereal rotation velocities of the sun. Solar Phys. 289, 1471–1476. DOI. ADS. Snodgrass, H.B., Dailey, S.B.: 1996, Meridional motions of magnetic features in the solar photosphere. Solar Phys. 163, 21. DOI. ADS. Snodgrass, H.B., Howard, R.: 1985, Torsional oscillations of low mode. Solar Phys. 95, 221. DOI. ADS. Stark, D., Wöhl, H.: 1981, On the solar rotation elements as determined from sunspot observations. Astron. Astrophys. 93, 241. ADS. Stix, M.: 1989, The Sun’s differential rotation. In: Klare, G. (ed.) Reviews in Modern Astronomy, Reviews in Modern Astronomy 2, 248. DOI. ADS. Sudar, D., Skokic,´ I., Ruždjak, D., Brajša, R., Wöhl, H.: 2014, Tracing sunspot groups to determine angular momentum transfer on the Sun. Mon. Not. Roy. Astron. Soc. 439, 2377. DOI. ADS. Sudar, D., Skokic,´ I., Brajša, R., Saar, S.H.: 2015, Steps towards a high precision solar rotation profile: results from SDO/AIA coronal bright point data. Astron. Astrophys. 575, A63. DOI. ADS. Sudar, D., Saar, S.H., Skokic,´ I., Poljanciˇ c´ Beljan, I., Brajša, R.: 2016, Meridional motions and Reynolds stress from SDO/AIA coronal bright points data. Astron. Astrophys. 587, A29. DOI. ADS. Tuominen, I., Virtanen, H.: 1984, Oscillatory motions in the sun corresponding to the solar cycle. Astron. Nachr. 305, 225. DOI. ADS. Ulrich, R.K., Boyden, J.E., Webster, L., Padilla, S.P., Snodgrass, H.B.: 1988, Solar rotation measurements at Mount Wilson. V – Reanalysis of 21 years of data. Solar Phys. 117, 291. DOI. ADS. Vršnak, B., Brajša, R., Wöhl, H., Ruždjak, V., Clette, F., Hochedez, J.-F.: 2003, Properties of the solar velocity field indicated by motions of coronal bright points. Astron. Astrophys. 404, 1117. DOI. ADS. Ward, F.: 1965, The general circulation of the solar atmosphere and the maintenance of the equatorial accel- eration. Astrophys. J. 141, 534. DOI. ADS. Ward, F.: 1966, Determination of the solar-rotation rate from the motion of identifiable features. Astrophys. J. 145, 416. DOI. ADS. Wayman, P.A.: 1980, Transactions of the International Astronomical Union, Volume XVIIB: Proceedings of the Seventeenth General Assembly, Montreal, 1979. ADS. Wöhl, H.: 1983, On the determination of heliographic positions and rotation velocities of sunspots. I – Com- parison of results from different observatories and different observing procedures. Solar Phys. 88, 65. DOI. ADS. Wöhl, H., Brajša, R.: 2001, Meridional motions of stable recurrent sunspot groups. Solar Phys. 198, 57. ADS. Wöhl, H., Brajša, R., Hanslmeier, A., Gissot, S.F.: 2010, A precise measurement of the solar differential rotation by tracing small bright coronal structures in SOHO-EIT images. Results and comparisons for the period 1998 – 2006. Astron. Astrophys. 520, A29. DOI. ADS. Zhao, J., Kosovichev, A.G.: 2004, Torsional oscillation, meridional flows, and vorticity inferred in the upper convection zone of the sun by time-distance helioseismology. Astrophys. J. 603, 776. DOI. ADS. Zuccarello, F., Zappalá, R.A.: 2003, Angular velocity during the cycle deduced using the sunspot group age selection methodology. Astron. Nachr. 324, 464. DOI. ADS. 5 Conclusion

One of the main goals of this PhD thesis was to investigate the surface solar differential rotation and the relationship between the solar rotation and activity (proposed in Paper I, II and III). The next goal was to provide a better understanding of the properties of the solar velocity field (rotational and meridional flows, their correlation and covariance, Reynolds stress) indicated by motions of the CBPs and sunspot groups in order to re- veal the transfer of angular momentum from higher to lower latitudes to maintain the observed differential rotation profile (proposed in Paper IV and V). The surface solar differential rotation was investigated in Paper I, by tracing sunspot groups during the 1964 - 2016 period and using the KSO sunspot drawings and white light images. There is one very important fact concerning the analysis presented in this study: until now, such analysis on KSO has been performed only for the years before 1980. The automatic method for determination of the heliographic coordinates on white light images showed reduced reliability for the years before 2009. Therefore, measure- ments for the 1964-2009 time period had to be done using the interactive procedure for coordinate determination. For the first time, the whole solar cycle no. 21, as well as the solar cycles nos. 22-24, were examined using the KSO sunspot drawings and white light images. Calculated rotation profiles derived from the KSO data and presented in this paper are in good agreement with DPD and GPR profiles derived in other stud- ies, which means that KSO provides a valuable data set with a satisfactory accuracy suitable for the investigation of the long term variabilities in the solar rotation profile and similar studies. This measurements have opened the way for new studies, as they provide a sample of 12152 sunspot groups which correspond to some 70000 individual positions of sunspot groups and cover 4 solar cycles. We made use of the KSO sunspot drawings and white light images and, therefore, continue our analysis of the KSO data,

88 by investigating the temporal variation of the solar rotation parameters (cycle to cycle variations, variations in single years, variations during the solar cycle) and the relation- ship between the parameters of the solar rotation and the solar activity using the yearly values of the solar rotation parameters. The KSO data has also never been used before for this purpose for the years after 1980, as well as for an analysis of meridional motions and horizontal Reynolds stress (Ruždjak et al. 2018). The dependence of the surface solar differential rotation on time and on the phase of the solar cycle, as well as the relationship between solar rotation and activity was in- vestigated in Paper II. Earlier observational findings, obtained by different tracers and methods, were confirmed: the solar rotation parameter A changes as the solar cycle evolves, showing lower value during the maximum of the solar cycle 23. The depen- dence of the parameter B on time and on the phase of the solar cycle was not confirmed. This is explained as a consequence of the determination of the solar rotation parameters by using data from all latitudes, even from polar regions, where the noise and various sources of systematic errors are much larger than at low latitudes. Consequently the variations of the parameter B are less noticeable. Another important conclusion of this study, which allows comparison of the results for sunspots and CBPs in future studies, is that the solar rotation parameter A determined by tracing CBPs represents a very good proxy for the solar equatorial rotation determined by sunspots and sunspot groups. This conclusion has been yielded from the comparison of the equatorial rotation velocity of CBPs, averaged over the whole observing period 1998−2006, and the equatorial rotation velocity of sunspots and sunspot groups provided in Table 16 of Wöhl et al. (2010). Investigation of the relationship between the solar rotation and activity (Paper II) is performed using the statistical analysis of the dependence between the annual and monthly values of the solar sidereal rotation parameters A and B and the solar activity (represented by SN and SSN). The statistically significant correlation was found for the solar rotation parameter A, whilst a very low and insignificant correlation was obtained for the rotation parameter B. From the dependence of the annual values of the solar side- real rotation parameters A and B and the interplanetary magnetic field strength, derived from the interdiurnal variability (IDV) index, another hypothesis may be adopted. It is about the corrrelation between the solar rotation and activity during the Maunder mini- mum, suposing that there should be obtained the strongest effect of this correlation. In Svalgaard & Cliver (2007, 2010) an expected value for the interplanetary magnetic field during the Maunder minimum is estimated (4 nT), while the value for the solar rotation parameter A for the Maunder minimum is suposed to be about 0.2 deg/day higher than average value for sunspots in usual solar activity conditions, around 14.7 deg/day (Eddy et al. 1976). The interplanetary magnetic field value of 4 nT corresponds exactly to the assumed value of the solar rotation parameter A, as obtained in the Figure 7 of the Paper

89 II. This conclusion may in some way be considered as a confirmation of an evidence that sunspots rotated faster at low latitudes with a more pronounced differential profile during the Maunder minimum (Eddy et al. 1976), e.g. that the equatorial solar rotation and activity are strongly correlated. This effect has also been discussed by Ruždjak et al. (2017) - Paper III. Meridional flows, torsional oscillations, horizontal Reynolds stress and conse- quently transport of the angular momentum towards the solar equator have been in- vestigated in Paper IV by tracing CBPs on SDO/AIA images, and in Paper V by tracing sunspot groups using the DPD data. The results for the meridional flow are the most controversial as they show a large number of possible orientations of the meridional flow, depending on the tracers or techniques used (Sec. 2.5). In Paper IV, a quote by Hathaway (1996) is given, as perhaps the most illustrative: “Unfortunately, previous measurements of the Sun’s meridional circulation have produced a bewildering array of results that appear to be of no help at all in constraining theory.” Calculations of the meridional flow in Paper IV and V also show different results: the motions towards the poles, for all latitudes, and the motions towards the zone of solar activity, respectively. We tried to consistently interpret these differing results in Paper V. The difference in measured meridional flow by sunspot groups in comparison to Doppler measurements and some other tracer measurements, e.g. CBPs, is interpreted as a consequence of dif- ferent flow patterns inside and outside of active regions. The different observations of the meridional motion of tracers, towards the centre of activity or oppositely, are sug- gested to be a consequence of the assumption that it is irrelevant to which latitude one assigns the observed meridional velocity (Olemskoy & Kitchatinov 2005). In both papers, Paper IV and V, the horizontal Reynolds stress showed transfer of angular momentum towards the solar equator. A statistically significant dependence of meridional motion on rotation velocity residuals obtained in Paper V also showed trans- fer of angular momentum towards the solar equator. The transport of angular momen- tum towards the solar equator, obtained by Reynolds stresses, conclusively describes the observed solar differential rotation profile, while the model of axisymmetric merid- ional circulation becomes quite disfavoured, as it requires surface meridional motions in the direction towards the solar equator, which is very rarely observed.

In summary, the study of the solar magnetic cycle, e.g. solar dynamo, that is thought to be hydromagnetic in nature, has been guided by the observations of the large scale, az- imuthally averaged, plasma motions, namely differential rotation and meridional flow. Thus, theoretical models have to deal with the explanation for both, the fluid motions and the magnetic field dynamics. The physical processes underlying this complex mag-

90 netohydrodynamical system are still not fully understood, and a large number of dynamo models have been considered until now, but still, there is not a model able to reproduce all the solar features. Therefore, precise measurements of the solar differential rota- tion, its variations, its correlation with the solar activity, as well as understanding of the properties of the solar velocity field (meridional motions, rotation velocity residuals, horizontal Reynolds stress) give important observational constraints on the modelling of the solar dynamo. That is why the analyses presented in this PhD thesis are important.

91 Bibliography

Badalyan, O. G. & Obridko, V. N. 2017, Monthly Notices of the Royal Astronomical Society, 466, 4535 Balthasar, H., Vázquez, M., & Wöhl, H. 1986, Astronomy and Astrophysics, 155, 87 Beck, J. G. 2000, Solar Physics, 191, 47 Beckers, J. M. 1981, in The Sun as a Star, ed. S. Jordan, NASA Conference Publication, 11 Belvedere, G., Godoli, G., Motta, S., Paterno, L., & Zappala, R. A. 1976, Solar Physics, 46, 23 Bertello, L., Henney, C. J., Ulrich, R. K., et al. 2000, The Astrophysical Journal, 535, 1066 Biermann, L. 1941, Vierteljahresschrift der Astronomischen Gesellschaft, 76, 194 Biermann, L. 1951, Zeitschrift für Astrophysik, 28, 304 Bogart, R. S. 1987, Solar Physics, 110, 23 Brajša, R., Ruždjak, D., & Wöhl, H. 2006, Solar Physics, 237, 365 Brajša, R., Wöhl, H., Vršnak, B., et al. 2004, Astronomy and Astrophysics, 414, 707 Brajša, R., Wöhl, H., Vršnak, B., et al. 2008, Central European Astrophysical Bulletin, 32, 165 Brun, A. S., Miesch, M. S., & Toomre, J. 2004, The Astrophysical Journal, 614, 1073 Brun, A. S. & Rempel, M. 2009, Space Science Reviews, 144, 151 Cameron, R. H., Dikpati, M., & Brandenburg, A. 2017, Space Science Reviews, 210, 367 Carrington, R. C. 1863, Observations of the Spots on the Sun from November 9, 1853, to March 24, 1861, made at Redhill (Williams and Norgate, London, Edinburgh) D’Silva, S. & Howard, R. F. 1995, Solar Physics, 159, 63 Durney, B. R. & Roxburgh, I. W. 1971, Solar Physics, 16, 3 Duvall, Jr., T. L. 1979, Solar Physics, 63, 3 Duvall, Jr., T. L., Dziembowski, W. A., Goode, P. R., et al. 1984, Nature, 310, 22 Eddy, J. A., Gilman, P. A., & Trotter, D. E. 1976, Solar Physics, 46, 3 Foukal, P. 1990, Solar astrophysics (New York, Wiley-Interscience) Foukal, P. 2013, Solar astrophysics (New York, Wiley-VCH,3rd edition) Gilman, P. A. 1981, NASA Special Publication, 450, 231 Gilman, P. A. 1983, in IAU Symposium, Vol. 102, Solar and Stellar Magnetic Fields: Origins and Coronal Effects, ed. J. O. Stenflo, 247–269 Gilman, P. A. & Howard, R. 1984a, Solar Physics, 93, 171 Gilman, P. A. & Howard, R. 1984b, The Astrophysical Journal, 283, 385 Gilman, P. A. & Miller, J. 1986, The Astrophysical Journal, Supplement, 61, 585 Glatzmaier, G. A. 1985, The Astrophysical Journal, 291, 300 Glatzmaier, G. A. 1987, in Astrophysics and Space Science Library, Vol. 137, The Internal Solar

92 Angular Velocity, ed. B. R. Durney & S. Sofia, 263 González Hernández, I., Howe, R., Komm, R., & Hill, F. 2010, Astrophysical Journal, Letters, 713, L16 González Hernández, I., Kholikov, S., Hill, F., Howe, R., & Komm, R. 2008, Solar Physics, 252, 235 Guerrero, G., Smolarkiewicz, P. K., Kosovichev, A. G., & Mansour, N. N. 2013, The Astro- physical Journal, 779, 176 Gupta, S. S., Sivaraman, K. R., & Howard, R. F. 1999, Solar Physics, 188, 225 Hathaway, D. H. 1996, Astrophysical Journal, 460, 1027 Hathaway, D. H. 2010, Living Reviews in Solar Physics, 7, 1 Howard, R. 1978, Reviews of Geophysics and Space Physics, 16, 721 Howard, R. 1979, Astrophysical Journal, Letters, 228, L45 Howard, R. 1984, Annual review of astronomy and astrophysics, 22, 131 Howard, R., Boyden, J. E., & Labonte, B. J. 1980, Solar Physics, 66, 167 Howard, R. & Gilman, P. A. 1986, Astrophysical Journal, 307, 389 Howard, R. F. 1991, Solar Physics, 135, 327 Hurlburt, N. E., Matthews, P. C., & Rucklidge, A. M. 2000, Solar Physics, 192, 109 Hurlburt, N. E. & Rucklidge, A. M. 2000, Monthly Notices of the Royal Astronomical Society, 314, 793 Jurdana-Šepić, R., Brajša, R., Wöhl, H., et al. 2011, Astronomy & Astrophysics, 534, A17 Kambry, M. A., Nishikawa, J., Sakurai, T., Ichimoto, K., & Hiei, E. 1991, Solar Physics, 132, 41 Karachik, N., Pevtsov, A. A., & Sattarov, I. 2006, The Astrophysical Journal, 642, 562 Kippenhahn, R. 1963, The Astrophysical Journal, 137, 664 Kitchatinov, L. L. 2005, Physics Uspekhi, 48, 449 Kitchatinov, L. L. 2011, in Astronomical Society of India Conference Series, Vol. 2, Solar dif- ferential rotation: origin, models and implications for dynamo, ed. R. S. P. Choudhuri, A. R. & D. Banerjee, 71 Kitchatinov, L. L. & Nepomnyashchikh, A. A. 2016, Advances in Space Research, 58, 1554 Kitchatinov, L. L. & Olemskoy, S. V. 2011, Monthly Notices of the Royal Astronomical Society, 411, 1059 Kitchatinov, L. L., Pipin, V. V., & Rüdiger, G. 1994, Astronomische Nachrichten, 315, 157 Kitchatinov, L. L. & Rüdiger, G. 1993, Astronomy and Astrophysics, 276, 96 Kitchatinov, L. L. & Rüdiger, G. 2005, Astronomische Nachrichten, 326, 379 Köhler, H. 1970, Solar Physics, 13, 3 Komm, R. W., Howard, R. F., & Harvey, J. W. 1993, Solar Physics, 145, 1 Kosovichev, A. G., Schou, J., Scherrer, P.H., et al. 1997, in IAU Symposium, Vol.181, Sounding Solar and Stellar Interiors, ed. J. Provost & F.-X. Schmider, 203 Kosovichev, A. G. & Zhao, J. 2016, in Lecture Notes in Physics, Vol. 914, Reconstruction of Solar Subsurfaces by Local Helioseismology, ed. J. P. Rozelot & C. Neiner, 25 Küker, M., Rüdiger, G., & Kitchatinov, L. L. 1993, Astronomy and Astrophysics, 279, L1

93 Labonte, B. J. & Howard, R. 1982, Solar Physics, 80, 361 Lanza, A. F. 2006, Monthly Notices of the Royal Astronomical Society, 373, 819 Lanza, A. F. 2007, Astronomy and Astrophysics, 471, 1011 Lebedinskii, A. I. 1941, Astronomicheskii Zhurnal, 18, 10 Leka, K. D. & Skumanich, A. 1998, The Astrophysical Journal, 507, 454 Li, K. J., Feng, W., Shi, X. J., et al. 2014, Solar Physics, 289, 759 Libbrecht, K. G. 1988, in ESA Special Publication, Vol. 286, Seismology of the Sun and Sun- Like Stars, ed. E. J. Rolfe, 131 Lustig, G. 1983, Astronomy and Astrophysics, 125, 355 Lustig, G. & Wöhl, H. 1990, Astronomy and Astrophysics, 229, 224 Lustig, G. & Wöhl, H. 1994, Solar Physics, 152, 221 Martinez Pillet, V. 1997, in Astronomical Society of the Pacific Conference Series, Vol. 118, Advances in Physics of Sunspots, ed. B. Schmieder, J. C. del Toro Iniesta, & M. Vazquez, 212 McIntosh, P. S. 1981, in The Physics of Sunspots, ed. L. E. Cram & J. H. Thomas, 7–54 McIntosh, S. W. & Gurman, J. B. 2005, Solar Physics, 228, 285 Meunier, N., Nesme-Ribes, E., & Collin, B. 1997, Astronomy and astrophysics, 319, 683 Miesch, M. S., Elliott, J. R., Toomre, J., et al. 2000, The Astrophysical Journal, 532, 593 Mulec, M. 2007, Master’s thesis, Institut for Physics, Karl-Franzens University Graz, Univer- sität Platz 3, Graz, Austria Nesme-Ribes, E., Ferreira, E. N., & Vince, L. 1993, Astronomy and Astrophysics, 276, 211 Newton, H. W. & Nunn, M. L. 1951, Monthly Notices of the Royal Astronomical Society, 111, 413 Olemskoy, S. V. & Kitchatinov, L. L. 2005, Astronomy Letters, 31, 706 Parker, E. N. 1979, The Astrophysical Journal, 230, 905 Perez Garde, M., Vazquez, M., Schwan, H., & Woehl, H. 1981, Astronomy and Astrophysics, 93, 67 Pidatella, R. M., Stix, M., Belvedere, G., & Paterno, L. 1986, Astronomy & Astrophysics, 156, 22 Pinsonneault, M. H., Kawaler, S. D., Sofia, S., & Demarque, P. 1989, The Astrophysical Journal, 338, 424 Pipin, V. V. & Kosovichev, A. G. 2018, The Astrophysical Journal, 854, 67 Poljančić, I., Brajša, R., Hržina, D., et al. 2011, Central European Astrophysical Bulletin, 35, 59 Poljančić Beljan, I., Jurdana-Šepić, R., Brajša, R., et al. 2017, Astronomy and Astrophysics, 606, A72 Poljančić Beljan, I., Jurdana-Šepić, R., Čargonja, M., et al. 2014, Central European Astrophysi- cal Bulletin, 38, 87 Poljančić, I., Brajša, R., Ruždjak, D., et al. 2010, Sun and Geosphere, 5, 52 Priest, E. R., Parnell, C. E., & Martin, S. F. 1994, The Astrophysical Journal, 427, 459 Pulkkinen, P. & Tuominen, I. 1998, Astronomy and Astrophysics, 332, 755 Rast, M. P., Ortiz, A., & Meisner, R. W. 2008, The Astrophysical Journal, 673, 1209 Rucklidge, A. M., Schmidt, H. U., & Weiss, N. O. 1995, Monthly Notices of the Royal Astro-

94 nomical Society, 273, 491 Rüdiger, G. 1980, Geophysical and Astrophysical Fluid Dynamics, 16, 239 Rüdiger, G. 1989, Differential rotation and stellar convection (Berlin: Akademie Verlag) Rüdiger, G., Egorov, P., & Ziegler, U. 2005, Astronomische Nachrichten, 326, 315 Rüdiger, G. & Hollerbach, R. 2004, The magnetic universe: geophysical and astrophysical dy- namo theory (Wiley-VCH) Rüdiger, G. & Kitchatinov, L. L. 1996, The Astrophysical Journal, 466, 1078 Rüdiger, G. & Tuominen, I. 1987, in Astrophysics and Space Science Library, Vol. 137, The Internal Solar Angular Velocity, ed. B. R. Durney & S. Sofia, 361–370 Ruždjak, D., Brajša, R., Sudar, D., Skokić, I., & Poljančić Beljan, I. 2017, Solar Physics, 292, 179 Ruždjak, D., Ruždjak, V., Brajša, R., & Wöhl, H. 2004, Solar Physics, 221, 225 Ruždjak, D., Sudar, D., Brajša, R., et al. 2018, Solar Physics, 293, 59 Sattarov, I., Pevtsov, A. A., Hojaev, A. S., & Sherdonov, C. T. 2002, The Astrophysical Journal, 564, 1042 Schatzman, E. 1959, in IAU Symposium, Vol. 10, The Hertzsprung-Russell Diagram, ed. J. L. Greenstein, 129 Scheiner, C. 1630, Rosa Ursina Schou, J., Antia, H. M., Basu, S., et al. 1998, The Astrophysical Journal, 505, 390 Schröter, E. H. 1985, Solar Physics, 100, 141 Schröter, E. H. & Wöhl, H. 1976, Solar Physics, 49, 19 Simon, G. W. & Weiss, N. O. 1970, Solar Physics, 13, 85 Snodgrass, H. B. 1984, Solar Physics, 94, 13 Snodgrass, H. B. & Dailey, S. B. 1996, Solar Physics, 163, 21 Snodgrass, H. B. & Howard, R. 1985, Solar Physics, 95, 221 Stix, M. 1989, in Reviews in Modern Astronomy, Vol. 2, The Sun’s Differential Rotation, ed. G. Klare, 248–266 Stix, M. 2002, The sun: an introduction (Berlin: Springer-Verlag, 2nd ed.) Sudar, D., Brajša, R., Skokić, I., Poljančić Beljan, I., & Wöhl, H. 2017, Solar Physics, 292, 86 Sudar, D., Saar, S. H., Skokić, I., Poljančić Beljan, I., & Brajša, R. 2016, Astronomy & Astro- physics, 587, A29 Sudar, D., Skokić, I., Ruždjak, D., Brajša, R., & Wöhl, H. 2014, Monthly Notices of the RAS, 439, 2377 Svalgaard, L. & Cliver, E. W. 2007, The Astrophysical Journal Letters, 661, L203 Svalgaard, L. & Cliver, E. W. 2010, Journal of Geophysical Research (Space Physics), 115, A09111 Tassoul, J.-L. & Tassoul, M. 2004, A concise history of solar and stellar physics (Princeton University Press) Thomas, J. H. & Weiss, N. O. 2008, Sunspots and Starspots (Cambridge University Press) Thomas, J. H., Weiss, N. O., Tobias, S. M., & Brummell, N. H. 2002, Nature, 420, 390 Tildesley, M. J. 2003, Monthly Notices of the Royal Astronomical Society, 338, 497

95 Tildesley, M. J. & Weiss, N. O. 2004, Monthly Notices of the Royal Astronomical Society, 350, 657 Tuominen, I. & Virtanen, H. 1984, Astronomische Nachrichten, 305, 225 Tuominen, J. 1961, Zeitschrift für Astrophysik, 51, 91 Tuominen, J. & Kyrolainen, J. 1982, Solar Physics, 79, 161 Usoskin, I. G. 2017, Living Reviews in Solar Physics, 14, 3 Vrabec, D. 1974, in IAU Symposium, Vol. 56, Chromospheric Fine Structure, ed. R. G. Athay, 201 Vršnak, B., Brajša, R., Wöhl, H., et al. 2003, Astronomy and Astrophysics, 404, 1117 Ward, F. 1965, The Astrophysical Journal, 141, 534 Weiss, N. O. 1965, The Observatory, 85, 37 Weiss, N. O., Thomas, J. H., Brummell, N. H., & Tobias, S. M. 2004, The Astrophysical Journal, 600, 1073 Wilson, P. R. 1968, Solar Physics, 3, 243 Wilson, P. R. 1994, Solar and stellar activity cycles (Cambridge, MA: Cambridge University Press, Cambridge), 25–45 Wilson, P. R., Burtonclay, D., & Li, Y. 1997, The Astrophysical Journal, 489, 395 Wöhl, H. & Brajša, R. 2001, Solar Physics, 198, 57 Wöhl, H., Brajša, R., Hanslmeier, A., & Gissot, S. F. 2010, Astronomy and Astrophysics, 520, A29 Zhao, J. & Kosovichev, A. G. 2004, Astrophysical Journal, 603, 776 Zuccarello, F. & Zappalá, R. A. 2003, Astronomische Nachrichten, 324, 464 Zwaan, C. 1978, Solar Physics, 60, 213 Zwaan, C. 1992, in NATO Advanced Science Institutes (ASI) Series C, Vol. 375, The evolution of sunspots, ed. J. H. Thomas & N. O. Weiss, 75–100

96