Skip to main content
Log in

Radial Anisotropy in the Upper Crust Beneath the Tehran Basin and Surrounding Regions

  • Published:
Pure and Applied Geophysics Aims and scope Submit manuscript

Abstract

Radial anisotropy is characterized by Rayleigh–Love wave discrepancy indicating crustal past and ongoing deformation. For this study, data from 1075 micro-earthquakes were collected based on recordings by 36 stations between 2004 and 2016, and after using data selection criteria, 375 events were used to calculate radial anisotropy in the Tehran region. More than 1908 and 1705 source-station Rayleigh and Love wave group velocity dispersion curves were, respectively, measured in the period band of 0.6–3.0 s. Furthermore, the tomographic inversion method was carried out to obtain group velocity maps for each period individually. Next, a damped least square iterative process was performed using a 3.5 × 3.5 km geographic grid size to calculate both VSV and VSH models. Horizontal and vertical spatial extents of the radial anisotropy beneath the Tehran region are revealed as maps of anisotropy as a percentage. Furthermore, the average value of the radial anisotropy as a function of depth indicates three sharp anomalies including: (1) relatively negative within a depth range from subsurface to 1.5 km, (2) relatively positive anomaly within a depth range from 2.0 to 4.0 km, and (3) an approximately isotropic half-space for a depth greater than 4.0 km. In general, the redeposition of former sediments near fault systems, geological and tectonic setting features are correlated with radial anisotropy anomalies at various depths as shown in the horizontal maps. In the radial anisotropy profiles, sedimentary thickness varies from ~ 500 to ~ 1500 m for southward transects, and it is constant for the eastward transect. These profiles clearly indicate the edges of three different tectonic settings.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

References

  • Abbassi, A., Nasrabadi, A., Tatar, M., Yaminifard, F., Abbassi, M. R., Hatzfeld, D., et al. (2010). Crustal velocity structure in the southern edge of the Central Alborz (Iran). Journal of Geodynamics, 49, 68–78.

    Article  Google Scholar 

  • Abbassi, M. R., & Shabanian, E. (1999). Fault map of North Tehran. Tehran: International Institute of Earthquake Engineering and Seismology.

    Google Scholar 

  • Ambraseys, N., & Melville, C. (1982). A history of persian earthquakes. Cambridge: Cambridge University Press.

    Google Scholar 

  • Ashtari, M., Hatzfeld, D., & Kamalian, N. (2005). Microseismicity in the region of Tehran. Tectonophysics, 395, 193–208.

    Article  Google Scholar 

  • Berberian, M. (1995). Master blind thrust faults hidden under the Zagros folds, active basement tectonics and surface morphotectonics. Tectonophysics, 241, 193–224.

    Article  Google Scholar 

  • Dziewonski, A., & Anderson, D. (1981). Preliminary reference Earth model. Physics of the Earth and Planetary Interiors, 25, 297–356.

    Article  Google Scholar 

  • Dziewonski, A., Bloch, S., & Landisman, M. (1969). A technique for the analysis of transient seismic signals. Bulletin of the Seismological Society of America, 59, 427–444.

    Google Scholar 

  • Englance, M. (1968) Geologie, Geomorphologie, hydrology de la Région de Tehran (Iran). Monpollier (France), p 139.

  • Gaherty, J. B., Wang, Y. B., Jordan, T. H., & Weidner, D. J. (1999). Testing plausible upper-mantle compositions using fine-scale models of the 410-km discontinuity. Geophysical Research Letters, 26, 1641–1644.

    Article  Google Scholar 

  • Guo, Z., Xing, G., Wang, W., & Yao, Z. (2012). Upper- and mid-crustal radial anisotropy beneath the central Himalaya and southern Tibet from seismic ambient noise tomography. Geophysical Journal International, 189, 1169–1182. https://doi.org/10.1111/j.1365-246X.2012.05425.x.

    Article  Google Scholar 

  • Herin, E., & Goforth, T. (1977). Phase-match filters: Application to study of Rayleigh waves. Bulletin of the Seismological Society of America, 67, 1259–1275.

    Google Scholar 

  • Herrmann, R. B., & Ammon, C. J. (2002). Computer programs in seismology-surface waves, receiver functions and crustal structure. Saint Louis University. http://www.eas.slu.edu/People/RBHerrmann/ComputerPrograms.html. Accessed Aug 2018.

  • Karato, S.-I., Jung, H., Katayama, I., & Skemer, P. (2008). Geodynamic significance of seismic anisotropy of the upper mantle: New insights from laboratory studies. Annual Review of Earth and Planetary Sciences, 36, 59–95.

    Article  Google Scholar 

  • Rawlinson, N. (2005). FMST: Fast Marching Surface Tomography Package, ACT 0200 (p. 0200). Canberra: Research School of Earth Sciences, Australian National University.

    Google Scholar 

  • Rieben, E. H. (1955). The geology of Tehran plain. American Journal of Science, 253, 617–639.

    Article  Google Scholar 

  • Rieben, E. H. (1966). Geological observation on alluvial deposits in northern Iran (p. 9). Tehran: Geological Survey of Iran.

    Google Scholar 

  • Sadidkhouy, A., Javan-Doloei, Gh, & Siahkoohi, H. R. (2008). Seismic anisotropy in the crust and upper mantle of the Central Alborz region, Iran. Tectonophysics, 456, 194–205.

    Article  Google Scholar 

  • Sadidkhouy, A., Javan-Doloei, Gh, & Siahkoohi, H. R. (2009). Mantle anisotropy in the Central Alborz obtained from SKS analysis, Iran. Journal of Geophysics, 2(2), 1–12.

    Google Scholar 

  • Shafiee, A., & Azadi, A. (2007). Shear-wave velocity characteristics of geological units throughout Tehran city, Iran. Journal of Asian Earth Sciences, 29, 105–115.

    Article  Google Scholar 

  • Shapiro, N. M., Ritzwoller, M. H., Molnar, P., & Levin, V. (2004). Thinning and flow of Tibetan crust constrained by seismic anisotropy. Science, 305(9), 233–236.

    Article  Google Scholar 

  • Shirzad, T., Naghavi, M., & Yamini Fard, F. (2018). Shallow/upper crustal shear wave structure of the Tehran region (Central Alborz, Iran) from the inversion of Rayleigh wave dispersion measurements. Journal of Seismology. https://doi.org/10.1007/s10950-018-9774-5.

    Google Scholar 

  • Shirzad, T., & Shomali, Z. H. (2013). Shallow crustal structures of the Tehran basin in Iran resolved by ambient noise tomography. Geophysical Journal International, 196, 1162–1176. https://doi.org/10.1093/gji/ggt449.

    Article  Google Scholar 

  • Shirzad, T., & Shomali, Z. H. (2014). Shallow crustal radial anisotropy beneath the Tehran Basin of Iran from seismic ambient noise tomography. Physics of the Earth and Planetary Interiors. https://doi.org/10.1016/j.pepi.2014.04.001.

    Google Scholar 

  • Shirzad, T., Shomali, Z. H., Riahi, M.-A., & Jarrahi, M. (2017). Near surface radial anisotropy in the Rigan Area/SE Iran. Tectonophysics, 694, 23–34. https://doi.org/10.1016/j.tecto.2016.11.036.

    Article  Google Scholar 

  • Silveira, G., & Stutzmann, E. (2002). Anisotropic tomography of the Atlantic Ocean. Physics of the Earth and Planetary Interiors, 132, 237–248.

    Article  Google Scholar 

  • Stein, S., & Wysession, M. (2003). An Introduction to seismology, earthquakes, and earth structure. Oxford: Blackwell.

    Google Scholar 

  • Stöcklin, J. (1974). Possible ancient continental margins in Iran. In C. A. Burk & C. L. Drake (Eds.), The geology of continental margins (pp. 873–887). New York: Springer.

    Chapter  Google Scholar 

  • Trampert, J., & Sneider, R. (1996). Model estimations biased by truncated expansions: Possible artifacts in seismic tomography. Science, 271, 1257–1260.

    Article  Google Scholar 

  • Vernant, P., Nilforoushan, F., Hatzfeld, D., Abbassi, M. R., Vigny, C., Masson, F., Nankali, H., Martinod, J., Ashtiani, A., Bayer, R., Tavakoli, F., & Chéry, J (2004) Contemporary crustal deformation and plate kinematics in Middle East constrained by GPS measurements in Iran and northern Oman. Geophysical Journal International, 157(1), 381–398.

    Article  Google Scholar 

  • Wessel, P., & Smith, W. H. F. (1998). New, improved version of generic mapping tools released. Eos, Transactions American Geophysical Union, 79, 579.

    Article  Google Scholar 

  • Xia, J., Miller, R. D., & Park, C. B. (1999). Estimation of near-surface shear wave velocity by inversion of Rayleigh waves. Geophysics, 64(3), 691–700.

    Article  Google Scholar 

  • Yanovskaya, T. B., Kizima, E. S., & Antonova, L. M. (1998). Structure of the crust in the Black Sea and adjoining regions from surface wave data. Journal of Seismology, 2, 303–316.

    Article  Google Scholar 

Download references

Acknowledgements

This work was supported by the Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP), Sao Paulo, Brazil [Grant numbers 2016/20952-4]. The digital micro earthquake dataset has been collected by the Tehran Disaster Mitigation and Management Organization (http://tdmmo.tehran.ir; not openly available to public; last accessed August 2018) and the Iranian Seismological Center (IrSC) at the University of Tehran/Iran (http://irsc.ut.ac.ir; not openly available to public; last accessed April 2018). All plots were also made using the Generic Mapping Tools (GMT) version 4 (Wessel and Smith 1998; http://www.soest.hawaii.edu/gmt, last accessed August 2018). We express our sincere thanks to IrSC, and TDMMO seismic networks for providing us with the data. We would also like to thank the Editor and two anonymous reviewers for their constructive comments and useful suggestions. A special note of thanks goes to Hassan Eliasi (from TDMMO) for digitizing faults and geological map.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to MohammadReza Hatami.

Appendices

Appendix 1

To investigate the checkerboard resolution test, an initial velocity model was used by adding ± 0.25 km/s to the observed mean value of the Rayleigh and Love group velocities at periods of 1, 2 and 3 s. The checkerboard cell size was also assigned three times the observed data inversion cell size (Trampert and Sneider 1996) and then synthetic travel times were individually calculated for all available source-station pairs in corresponding periods. As shown in Fig. 7, the checkerboard test resolutions were recovered reasonably well except for the southwestern part of the studied area. Unreliable regions, which are not recovered in the checkerboard test resolution in different periods, were masked. Thus, the region with good resolution for horizontal slices, vertical profiles and further interpretation were used. The corner of the input model is indicated by green addition symbols (+) on all recovered result maps.

Fig. 7
figure 7

Checkerboard test resolutions of Rayleigh (top maps) and Love (bottom maps) waves with periods of 1, 2 and 3 s

Appendix 2

Sensitivities of dispersion data are the most important issues in surface wave studies, especially in considering the discrepancy between fundamental mode of Rayleigh and Love waves based on a linearized inversion method. Short periods of surface waves carry information on relatively shallow structures than the longer periods. In other words, different periods of fundamental modes of Rayleigh and Love waves sample different depths of the velocity structure. Thus, the sensitivity kernel was calculated to obtain the maximum penetration depth and effective depth range using the average of local Rayleigh and Love wave dispersion curves. Figure 8 (1–6) show the normalized depth sensitivity kernel for the periods of 0.6, 1.0, 1.5, 2.0, 2.5 and 3.0 s, individually. Minima and maxima values of penetration depths are approximately 200 m and 6 km corresponding to 0.6 and 3 s, as shown in Fig. 8 (1, 6) respectively. However, surface wave has very small sensitivity around the maximum penetration depth. Thus, the effective depth sensitivity kernels were used for further interpretations. Moreover, differences in sensitivity kernels of surface wave as a function of depth may cause appear artifact anomalies in radial anisotropy result. Thus, to obtain different sensitivity kernels of Rayleigh and Love waves at period of 0.6 and 3.0 s, a similar formula [2 × [SenLove − SenRayleigh]/[SenLove + SenRayleigh]] was used. In this formula, SenRayleigh and SenLove are the sensitivity kernels of Rayleigh and Love waves as a function of depth respectively. Also, if SenRayleigh or SenLove closes to zero, the value of formula limits to ± 2%. As shown in Fig. 8 (7), the value of formula for minimum and maximum periods is between ± 1% in this study.

Fig. 8
figure 8

Normalized sensitivity kernel of Rayleigh (black) and Love (gray) waves with periods of 0.6 (1), 1.0 (2), 1.5 (3), 2.0 (4), 2.5 (5) and 3.0 (6) s. The effective depth sensitivity ranges of these periods are 0.0–0.7, 0.5–1.2, 0.9–2.5, 1.2–3.2, 1.5–4.0 and 1.8–5.0 km, respectively. The difference between the Rayleigh and Love wave sensitivity kernels, which is calculated using [2 × [SenLove − SenRayleigh]/[SenLove + SenRayleigh]] formula, are depicted by gray (0.6 s) and black (3 s) traces in (7)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Naghavi, M., Hatami, M., Shirzad, T. et al. Radial Anisotropy in the Upper Crust Beneath the Tehran Basin and Surrounding Regions. Pure Appl. Geophys. 176, 787–800 (2019). https://doi.org/10.1007/s00024-018-1986-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00024-018-1986-7

Keywords

Navigation