Log in

Constraints on Jalisco Block Motion and Tectonics of the Guadalajara Triple Junction from 1998–2001 Campaign GPS Data

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

Abstract

A GPS campaign network in the state of Jalisco was occupied for ~36 h per station most years between 1995 and 2005; we use data from 1998–2001 to investigate tectonic motion and interseismic deformation in the Jalisco area with respect to the North America plate. The twelve stations used in this analysis provide coverage of the Jalisco Block and adjacent North America plate, and show a pattern of motion that implies some contribution to Jalisco Block boundary deformation from both tectonic motion and interseismic deformation due to the offshore 1995 earthquake. The consistent direction and magnitude of station motion on the Jalisco Block with respect to the North America reference frame, ~2 mm/year to the southwest (95% confidence level), perhaps can be attributed to tectonic motion. However, some station velocities within and across the boundaries of the Jalisco Block are also non-zero (95% confidence level), and the overall pattern of station velocities indicates both viscoelastic response to the 1995 earthquake and partial coupling of the subduction interface (together termed “interseismic deformation”). Our results show motion across the northern Colima rift, the eastern boundary of the Jalisco Block, which is likely to be sinistral oblique extension rather than pure extension. We constrain extension across both the Colima rift and the northeastern boundary of the Jalisco Block, the Tepic-Zacoalco rift, to ≤8 mm/year (95% confidence level), slow compared to relative rates of motion at nearby plate boundaries.

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

Access this article

Price includes VAT (Brazil)

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

References

  • Allan, J. F. (1986), Geology of the Northern Colima and Zacoalco Grabens, southwest Mexico: Late Cenozoic rifting in the Mexican Volcanic Belt, Geol. Soc. of Am. Bull. 97, 473–485.

  • Allan, J. F., S. A Nelson, J. F. Luhr, I. S. E. Carmichael, M. Wopat, and P. J. Wallace (1991), Pliocene-Holocene Rifting and Associated Volcanism in Southwest Mexico: An Exotic Terrane in the Making, in Dauphin, J. P. & Simoneit, B. R. T., eds., The Gulf and Peninsular Province of the Californias: Boulder, Colorado, AAPG Memoir 47, 425–445.

  • Altamimi, Z., P. Sillard, and C. Boucher (2002), ITRF2000: A new release of the International Terrestrial Reference Frame for earth science applications, J. of Geophys. Res., 107(B10), 2214, doi:10.1029/2001JB000561.

  • Andrews, V., J. M. Stock, G. Reyes-Dávila and C. Ramírez-Vazquez, Double-difference relocation of the aftershocks of the Tecomán, Colima, Mexico earthquake of 22 January 2003, PAGEOPH, in press.

  • Arzate, J. A., R. Alvarez, V. Yutsis, J. Pacheco, and H. Lopez-Loera (2006), Geophysical modeling of Valle de Banderas Graben and its structural relation to Bahia de Banderas, Mexico, Revista Mexicana de Ciencias Geologicas 23 (2), 184–198.

  • Bandy, W. and M. Pardo (1994), Statistical examination of the existence and relative motion of the Jalisco and Southern Mexico Blocks, Tectonics 13 (4), 755–768.

  • Bandy, B. L., F. Michaud, J. Bourgois, T. Calmus, J. Dyment, C. A. Mortera-Gutierrez, J. Ortega-Ramirez, B. Pontoise, J.-Y. Royer, B. Sichler, M. Sosson, M. Rebolledo-Vieyra, F. Bigot-Cormier, O. Diaz-Molina, A. D. Hurtado-Artunduaga, G. Pardo-Castro, C. Trouillard-Perrot (2005), Subsidence and strike-slip tectonism of the upper continental slope off Manzanillo, Mexico, Tectonophysics 398, 115–140.

  • Becker, J. J. and D. T. Sandwell (2006), SRTM30_PLUS V2.0: Data fusion of SRTM land topography with measured and estimated seafloor topography, 29 July 2006, ftp://edcsgs9.cr.usgs.gov/pub/data/srtm/SRTM30.

  • Bourgois, J., V. Renard, J. Aubouin, W. Bandy, E. Barrier, T. Calmus, J.-L. Carfantan, J. Guerrero, J. Mammerickx, B. Mercier de Lepinay, F. Michaud, M. Sosson (1988), Active fragmentation of the North American Plate: offshore boundary of the Jalisco Block off Manzanillo, C. R. Acad., Sci. Paris 307 (II), 1121–1130.

  • Correa-Mora, F., C. DeMets, E. Cabral-Cano, B. Marquez-Azua, and O. Diaz-Molina (2008), Interplate coupling and transient slip along the subduction interface beneath Oaxaca, Mexico, Geophys. J. Int. 175, 269–290.

    Google Scholar 

  • DeMets, C. and D. S. Wilson (1997), Relative motions of the Pacific, Rivera, North American, and Cocos plates since 0.78 Ma, J. of Geophys. Res. 102 (B2), 2789–2896.

  • DeMets, C. and S. Traylen (2000), Motion of the Rivera plate since 10 Ma relative to the Pacific and North American plates and the mantle, Tectonophysics 318, 119–159.

  • Dong, D., T. A. Herring, and R. W. King (1998), Estimating regional deformation from a combination of space and terrestrial geodetic data, J. of Geodesy 72, 200 –214.

  • Feigl, K. L., D. C. Agnew, Y. Bock, D. Dong, A. Donnellan, B. H. Hager, T. A. Herring, D. D. Jackson, T. H. Jordan, R. W. King, S. Larsen, K. M. Larson, M. H. Murray, Z. Shen, and F. H. Webb (1993), Space Geodetic Measurement of Crustal Deformation in Central and Southern California, 1984–1992, J. of Geophys. Res. 98 (B12), 21,677–21,712.

  • Ferrari, L. (1995), Miocene shearing along the northern boundary of the Jalisco block and the opening of the southern Gulf of California, Geology 23 (8), 751–754.

  • Ferrari, L. (2004), Slab detachment control on mafic volcanic pulse and mantle heterogeneity in central Mexico, Geology 32 (1), 77–80.

  • Ferrari, L. and J. Rosas-Elguera (2000), Late Miocene to Quaternary extension at the northern boundary of the Jalisco block, western Mexico: The Tepic-Zacoalco rift revised, in Delgado-Granados, H., G. Aguirre, and J. M. Stock, eds., Cenozioc Tectonics and Volcanism of Mexico: Boulder, Colorado, GSA Special Paper 334, 41–63.

  • Ferrari, L., G. Pasquare, S. Venegas, D. Castillo, and F. Romero (1994), Regional tectonics of western Mexico and its implications for the northern boundary of the Jalisco Block, Geofisica Internacional 33 (1), 139–151.

  • Frey, H. M., R. A. Lange, C. M. Hall, H. Delgado-Granados, I. S. E. Carmichael (2007), A Pliocene ignimbrite flare-up along the Tepic-Zacoalco rift: Evidence for the initial stages of rifting between the Jalisco block (Mexico) and North America, GSA Bulletin 119 (1/2), 49–64.

  • Garduño-Monroy, V. H., R. Saucedo-Girón, Z. Jiménez, J. C. Gavilanes-Ruiz, A. Cortés-Cortés, and R. M. Uribe-Cifuentes (1998), La falla Tamazula, Límite suroriental del Bloque Jalisco, y sus relaciones con el complejo volcánico de Colima, México, Rev. Mex. Cienc. Geol. 15, 132–144.

  • Grand, S. P., T. Yang, S. Suhardja, D. Wilson, M. G. Speziale, J. Gomez Gonzalez, G. Leon-Soto, J. Ni, and T. Dominguez Reyes (2007), Seismic structure of the Rivera subduction zone – the MARS experiment (abstract), Eos, Trans., AGU, T32A-02.

  • Herring, T. A., J. L. Davis, and I. I. Shapiro (1990), Geodesy by Radio Interferometry: The Application of Kalman Filtering to the Analysis of Very Long Baseline Interferometry Data, J. of Geophys. Res. 95 (B8), 12,561–12,581.

  • Hutton, W., C. DeMets, O. Sanchez, G. Suarez, and J. Stock (2001), Slip kinematics and dynamics during and after the 1995 October 9 M w  = 8.0 Colima-Jalisco earthquake, Mexico, from GPS geodetic constraints, Geophys. J. Int. 146, 637–658.

  • Johnson, C. A. and C. G. A. Harrison (1990), Neotectonics in central Mexico, Physics of the Earth and Planetary Interiors 64, 187–210.

  • Khutorskoy, M. D., L. A. Delgado-Argote, R. Fernandez, V. I. Kononov, B. G. Polyak (1994), Tectonics of the offshore Manzanillo and Tecpan basins, Mexican Pacific, from heat flow, bathymetric and seismic data, Geofis. Int. 33, 161–185.

  • Kostoglodov, V. and W. Bandy (1995), Seismotectonic constraints on the convergence rate between the Rivera and North American plates, J. of Geophys. Res. 100 (B9), 17,977–17,989.

  • Luhr, J. F., S. A. Nelson, J. F. Allan, and I. S. E. Carmichael (1985), Active rifting in southwestern Mexico: Manifestations of an incipient eastward spreading-ridge jump, Geology 13, 54–57.

  • Mao, A., C. G. A. Harrison, and T. H. Dixon (1999), Noise in GPS coordinate time series, J. of Geophys. Res. 104 (B2), 2797–2816.

  • Marquez-Azua, B. M. and C. DeMets (2003), Crustal velocity field of Mexico from continuous GPS measurements, 1993 to June 2001: Implications for the neotectonics of Mexico, J. of Geophys. Res. 108 (B9), 2450, doi:10.1029/2002/2002JB002241.

  • Marquez-Azua, B. M. and C. DeMets (2009), Deformation of Mexico from continuous GPS from 1993 to 2008, Geochem. Geophys. Geosyst. 10, Q02003, doi:10.1029/2008GC002278.

  • Marquez-Azua, B. M., C. DeMets, and T. Masterlark (2002), Strong interseismic coupling, fault afterslip, and viscoelastic flow before and after the Oct. 9, 1995 Colima-Jalisco earthquake: Continuous GPS measurements from Colima, Mexico; Geophysical Research Letters 29 (N.8), 122-1–122-4.

  • Masterlark, T., C. DeMets, H. F. Wang, J. Stock, and O. Sanchez (2001), Homogeneous vs. heterogeneous subduction zone models: Coseismic and postseismic deformation, Geophysical Review Letters 28, 4047–4050.

  • Melbourne, T., I. Carmichael, C. DeMets, K. Hudnut, O. Sanchez, J. Stock, G. Suarez, F. Webb (1997), The geodetic signature of the M8.0 Oct. 9, 1995, Jalisco subduction earthquake, Geophys. Res. Let. 24 (6), 715–718.

  • Melbourne, T., F. Webb, J. Stock, and C. Reigber (2002), Rapid postseismic transients in subduction zones from continuous GPS, J. Geophys. Res. 107 (B10), 2241, doi:10.1029/20001JB000555.

  • Molnar, P. and J. M. Stock (1985), A method for bounding uncertainties in combined plate reconstructions, J. of Geophys. Res. 90 (B14), 13537–12544.

  • Nixon, G. T. (1982), The relationship between Quaternary volcanism in central Mexico and the seismicity and structure of subducted ocean lithosphere, Geol. Soc. of Am. Bull. 93, 514–523.

  • Nuñez-Cornú, F. J., R. L. Marta, F. A. Nava-P., G. Reyes-Davila, and C. Suarez-Plascencia (2002), Characteristics of seismicity in the coast and north of Jalisco Block, Mexico, Phys. of the Earth and Planet. Int. 132, 141-155.

    Google Scholar 

  • Pacheco, J., S. K. Singh, J. Dominguez, A. Hurtado, L. Quintanar, Z. Jimenez, J. Yamamoto, C. Gutierrez, M. Santoyo, W. Bandy, M. Guzman, V. Kostoglodov (1997), The October 9, 1995 Colima-Jalisco, Mexico earthquake (M W 8): An aftershock study and a comparison of this earthquake with those of 1932, Geophys. Res. Let. 24 (17), 2223–2226.

  • Pacheco, J. F., C. A. Mortera-Gutierrez, H. Delgado, S. K. Singh, R. W. Valenzuela, N. M. Shapiro, M. A. Santoyo, A. Hurtado, R. Barron, and E. Gutierrez-Moguel (1999), Tectonic significance of an earthquake sequence in the Zacoalco half-graben, Jalisco, Mexico, Journal of South American Earth Sciences 12, 557–565.

  • Pacheco, J., W. Bandy, G. A. Reyes-Davila, F. J. Nunez-Cornu, C. A. Ramirez-Vazquez, J. R. Barron (2003), The Colima, Mexico, earthquake (Mw 5.3) of 7 March 2000: seismic activity along the southern Colima rift, Bull. Seismol. Soc. Am. 93, 1458–1476.

  • Pardo, M. and G. Suarez (1995), Shape of the subducted Rivera and Cocos plates in southern Mexico: Seismic and tectonic implications, J. of Geophys. Res. 100 (B7), 12,357–12,373.

  • Rosas-Elguera, J., L. Ferrari, V. H. Garduño-Monroy, J. Urrutia-Fucugauchi (1996), Continental boundaries of the Jalisco block and their influence in the Pliocene-Quaternary kinematics of western Mexico, Geology 24 (10), 921–924.

  • Rosas-Elguera, J., L. Ferrari, M. Martinez, J. Urritia-Fucagauchi (1997), Stratigraphy and Tectonics of the Guadalajara Region and Triple-Junction Area, Western Mexico, International Geology Review 39, 125–140.

  • Schmitt, S. V., C. DeMets, J. Stock, O. Sanchez, B. Marquez-Azua, and G. Reyes (2007), A geodetic study of the 2003 January 22 Tecomán, Colima, Mexico earthquake, Geophys. J. Int. 169, 389–406.

  • Singh, S. K., L. Ponce, S. P. Nishenko (1985), The great Jalisco, Mexico, earthquakes of 1932: Subduction of the Rivera plate, Bull. Seis. Soc. Am. 75 (5), 1301–1313.

  • Stock, J. M. (1993), Tectónica de placas y la evolución del bloque Jalisco, México, GEOS, Bol. Mex Geofis. Union 13 (3), 3–9.

  • Suarez, G., V. Garcia-Acosta, and R. Gaulon (1994), Active crustal deformation in the Jalisco block, Mexico: evidence for a great historical earthquake in the 16 th century, Tectonophysics, 234, 117–127.

  • Suhardja, S., S. Grand, D. Wilson, M. Guzman-Speziale, J. Gomez-Gonzalez, J. Ni, and T. Dominguez-Reyes (2007), Crustal structure beneath Southwestern Mexico (abstract), Eos, Trans., AGU, S33A-1052.

  • Urrutia-Fucugauchi, J. and T. Gonzalez-Moran (2006), Structural pattern at the northwestern sector of the Tepic-Zacoalco rift and tectonic implications for the Jalisco block, western Mexico, Earth Planets Space 58, 1303–1308.

    Google Scholar 

  • Wang, K. (2007), Elastic and viscoelastic models of crustal deformation in subduction earthquake cycles, in The Seismogenic Zone of Subduction Thrust Faults, eds. T. H. Dixon and J. C. Moore, 540–575.

  • Wang, X., F. Niu, J. Ni, and S. Grand (2008), Crustal and Mantle Structure of the Jalisco Block of western Mexico from Surface Wave Tomography (abstract), Eos, Trans., AGU, S23A-1871.

  • Yagi, Y., T. Mikumo, J. Pachecho, and G. Reyes (2004), Source Rupture Process of the Tecomán, Colima, Mexico Earthquake of 22 January 2003, Determined by Joint Inversion of Teleseismic Body-Wave and Near-Source Data, Bull. Seism. Soc. Am. 94, 1795–1807.

  • Yang, T., S. P. Grand, D. Wilson, M. Guzman-Speziale, J. M. Gomez-Gonzalez, T. Dominguez-Reyes, and J. Ni (2009), Seismic structure beneath the Rivera subduction zone from finite-frequency seismic tomography, J. Geophys. Res. 114, B01302, doi:10.1029/2008JB005830.

  • Zumberge, J. F., M. B. Heflin, D. C. Jefferson, M. M. Watkins, and F. H. Webb (1997), Precise point positioning for the efficient and robust analysis of GPS data from large networks, J. of Geophys. Res. 102 (B3), 5005–5017.

Download references

Acknowledgments

We would like to thank Jeff Genrich and Tom Herring for helpful discussions. Support for this research was provided by National Science Foundation grants EAR-0510395 (J. Stock) and EAR-0510553 (C. DeMets).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michelle M. Selvans.

Appendices

Appendix 1: Methods

Data are processed using the software packages GAMIT and GLOBK, developed at the Massachusetts Institute of Technology by T. A. Herring and D. Dong (Herring et al., 1990; Feigl et al., 1993 ; Zumberge et al., 1997; Dong et al., 1998). Inputs into GAMIT are the data files, session specifications (year and day, receiver and antenna types, and antenna heights), and good initial station coordinates. Outputs are the loosely constrained solution files, which are passed to GLOBK for multi-session processing. A least-squares analysis is used to obtain the GAMIT solution files, and a combination of well-defined reference frame and Kalman filter with white noise are used in GLOBK to obtain the station coordinate time series and velocities.

The recursive, time-domain Kalman filter (run eight times) estimates the state of a dynamic system from a series of incomplete and noisy measurements, such as campaign GPS data (Mao et al., 1999); only the previous time step and the current measurement are needed to estimate the current state, with a linear relation used in the calculation, making it computationally efficient (Herring et al., 1990). The Kalman filter uses a multivariate normal distribution for the process noise, which is independent of past process noise for every time step (i.e. the Markov process). In the simplest case, the Markov process allows separate noise levels for the north, east, and vertical components of position (we use 2 mm/year in the horizontal directions, 5 mm/year in the vertical direction).

The same a priori position and velocity constraints for each station are used in GAMIT and GLOBK, and are obtained from standard ITRF2000 files for the North America reference stations, and from the online Scripps Coordinate Update Tool for the campaign stations. Rotation and translation of three components of position and their rates are permitted when determining the reference frame in GLOBK. We iterate the reference frame solution eight times in order to stabilize the coordinate system, with 75% weighting on the coordinate sigmas of the previous iteration, and a 4-sigma cutoff for sites that are discordant with a priori values. Height residuals allowed in the stabilization are limited to 5 mm between the best and median for position (and 5 mm/year for the related rate), and 3 mm for the rms position (and 3 mm/year for the related rate).

Earth orientation parameters are tightly constrained in GLOBK by Markov process values of 0.25 mas/day in orientation and 0.1 mas/day in its rate of change. Since final orbits are used, a priori GPS satellite orbital parameters are also tightly constrained, with correspondingly tightly constrained random walk variation allowed while processing multiple sessions. Changes to orbital parameters due to random noise are constrained to 10 cm/day in XYZ, 0.01 mm/s/day for the XYZ time derivatives, 1%/day in direct and y-bias non-gravitational parameters, 0.1%/day in b axis bias and once-per-rev parameters, and 1 cm/day for SV antenna offsets.

Appendix 2: Results

See Tables 1 and 2.

Table 1 Global [North America (NA)] and campaign GPS sites, their velocity components and 1σ errors (relative to North America, as defined by the stations with a *), and the cross-correlation (ρ) between north (N) and east (E) rates
Table 2 GPS sites in the Jalisco region, their velocity components and 1σ errors (calculated relative to North America, as defined in Table 1, and presented relative to TAPA, a campaign site on the Jalisco Block), and the cross-correlation (ρ) between N and E rates

Appendix 3: Assumptions used in triple junction constraints

We assume a simplified geometry: three blocks (the Jalisco Block (JB), North America (NA), and the Michoacan Block (MB) (e.g. Johnson and Harrison, 1990) meet at a continental triple junction formed by the Tepic–Zacoalco rift, the northern Colima rift, and the Chapala rift. GPS sites UGEO and UMON are on NA; GPS sites CEBO and CGUZ are on the MB; and GPS site TAPA is on JB. We assume a flat-earth geometry because of the close spacing of these stations (<100 km separation). We use the results from Appendix 2, Table 2 to constrain the velocity of NA and the MB relative to the JB, assuming no compression across any of the boundaries, as follows.

  1. 1.

    UGEO and UMON lie on NA, and should move together with respect to TAPA. Thus, the velocity of NA must lie within the intersection of the 95% confidence limits of the UMON and UGEO velocities relative to TAPA (red region in velocity diagram in Fig. 2b). Similarly, the velocity of the MB must lie within the intersection of the 95% confidence regions of the CEBO and CGUZ velocities.

  2. 2.

    We assume no compression across the Tepic–Zacoalco rift, which trends N56°W. Therefore, the velocity of NA with respect to TAPA must lie northeast of a line with an azimuth of N56°W. This confines the allowable velocities for NA to points within the region shown in red in Fig. 2c.

  3. 3.

    We assume no compression in the northern Colima rift, which trends N–S. Thus, the velocities of stations on the MB (CEBO and CGUZ) must lie east of a line trending N–S from the origin. This requires their velocities to lie within the green region shown in Fig. 2d.

  4. 4.

    We assume no compression across the east–west trending Chapala rift. This eliminates velocities of NA from Fig. 2c that lie south of the southernmost point in the green velocity field in Fig. 2d, yielding the possible velocities of NA relative to block JB, shown in red in Fig. 2e. Similarly, we eliminate velocities of stations CGUZ and CEBO from Fig. 2d that lie north of an E-W line that passes through the northernmost point of the allowed velocity field for NA in Fig. 2e. The velocity for the MB then is restricted to the green field of Fig. 2f.

This yields the following constraints on the velocity triangle at the Guadalajara triple junction. At 95% confidence, the velocity of NA relative to the JB can lie anywhere in the red shaded region of Fig. 2e. The velocity of the MB relative to the JB can lie anywhere in the green-shaded region on Fig. 2f. However, the combination of velocities (one point from the red field and one point from the green field) must further satisfy two additional constraints. First, the red point cannot lie south of the green one (otherwise there would be compression across the Chapala rift). Second, the points on the velocity triangle must have the same topology as the blocks in map view (i.e. the JB, NA, and the MB must be encountered in clockwise order going around the triangle). A velocity triangle with the JB, NA, and the MB in counterclockwise order would imply that at least one of the boundaries is compressional.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Selvans, M.M., Stock, J.M., DeMets, C. et al. Constraints on Jalisco Block Motion and Tectonics of the Guadalajara Triple Junction from 1998–2001 Campaign GPS Data. Pure Appl. Geophys. 168, 1435–1447 (2011). https://doi.org/10.1007/s00024-010-0201-2

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00024-010-0201-2

Keywords

Navigation