Geophys. J. Int. (1999) 139, 563^572 Length of day variations due to zonal tides for an inelastic earth in non-hydrostatic equilibrium P. Defraigne1 and I. Smits2 1 2 Royal Observatory of Belgium, Av. Circulaire 3, B-1180 Brussels, Belgium. E-mail: [email protected] Universitë Catholique de Louvain, 1348 Louvain-la-Neuve, Belgium Accepted 1999 July 12. Received 1999 June 21; in original form 1998 November 2 S U M M A RY We present a re¢ned theoretical model for length-of-day (lod) variations induced by the zonal part of the tide-generating potential. The model is computed from a numerical integration, from the Earth's centre up to the surface, of the equation of motion, the rheological equation of state and Poisson's equation. The Earth is modelled as a threelayered body, with an inelastic inner core, an inviscid £uid core and an inelastic mantle sustaining convection, which induces deviations from hydrostatic equilibrium. The model also incorporates ocean corrections deduced from dynamic ocean models. It is shown that the non-hydrostatic structure inside the Earth has an e¡ect of less than 0.1 per cent on the transfer functions, while the di¡erent modellings of mantle inelasticity (di¡erent combinations of possible values for the inelastic parameters) can lead to a wide range of results. Finally, we show that the precision of the geodetic observations of UT1 and the precision of the oceanic and atmospheric corrections are not yet su¤cient to obtain information about mantle inelasticity from the comparison between theoretical models and geodetic observations. Key words: anelasticity, length of day, zonal tides. 1 IN T ROD U C T I O N Geodetic observations during the last few decades have allowed the detection of £uctuations of the Earth's rotation speed. A part of these £uctuations comes from the response of the Earth to the zonal part of the luni-solar (and planetary) gravitational potential. Indeed, the periodic change of the principal moments of inertia associated with the tidal deformation induces, due to conservation of angular momentum, a periodic change in rotation rate, and hence a periodic change in the length of the day (lod). The principal periods of these variations correspond to 13.66 days, semi-annual, annual, monthly, trimonthly periods, etc. The amplitudes of these variations are of the order of 0.1 ms for the lod, which leads, for example, to about 4 ms for UT1 (the total rotational angle) for the semi-annual tide (Ssa). Such amplitudes are well above the precision of observational techniques such as VLBI, for which the accuracy is about 0.01 ms. From a theoretical point of view, Merriam (1980) has computed the zonal response coe¤cient i (as de¢ned by Agnew & Farrell 1978, and representing the transfer function for the lod changes induced by the tides) from the change in the moment of inertia for an earth without core^mantle coupling and with an equilibrium ocean tide. Merriam showed that, at periods from 13.66 days up to 1 year, a correspondence between the ß 1999 RAS computed and observed values of i exists only if the core and mantle are not coupled (free-slip condition). This implies that only variations of the polar moment of inertia C of the mantle and of the crust enter in the variations of lod. Yoder et al. (1981) computed the periodic variations of the Earth's rotation speed due to long-period tides directly from the ephemerides of the Moon and the Sun, and also pointed out the importance of decoupling between the core and the mantle in this kind of computation. These authors presented the results in the form of a harmonic series involving the di¡erent tidal frequencies scaled by a factor k/C, where k is the tidal e¡ective (i.e. considered only for crust and mantle) gravitational Love number and C is the polar moment of inertia of the crust and mantle. Wahr et al. (1981) solved the equations of in¢nitesimal elastic gravitational deformation of a rotating and slightly elliptical earth in response to an external gravitational potential. This method provides us with a transfer function that can be linked with the zonal response coe¤cient i. In such a computation, the core can slip freely within the mantle (no friction) so that the change of rotation rate is only due to variations of the inertia tensor of the mantle. In that paper, Wahr et al. also showed that additional e¡ects of viscous coupling at the core^ mantle boundary could induce out-of-phase perturbations on UT1 as large as 10 ms for the 18.6 yr tide. Nevertheless, this 563 564 P. Defraigne and I. Smits could not be extracted from the observations because of the poor understanding of the decade £uctuation associated with core^mantle interactions (which can be of the order of 1 s for timescales of about 20 yr). Further studies (Wahr & Bergen 1986; Merriam 1985) pointed out the e¡ect of mantle inelasticity on the lod variations induced by long-period tides. Their results show a frequency-dependent transfer function, while the transfer function is constant in the elastic case. This frequency dependence can be used to constrain mantle inelasticity from the observed lod variations, as will be discussed in this paper. More recently, the Earth's response to long-period tides has been determined in several studies by computing the factor k/C from di¡erent kinds of observations. Hefty & Capitaine (1990) analysed optical astrometry, VLBI and SLR data. Robertson et al. (1994) analysed only VLBI data. Schastok et al. (1994) used VLBI and SLR observations. Another set of observed values was provided by McCarthy & Luzum (1993) from VLBI, SLR and LLR data. Nam & Dickman (1990; see also Dickman & Nam 1995) computed the zonal response coe¤cient i deduced from VLBI (IRIS) observations after correcting the data for atmospheric wind and pressure e¡ects. They also used a dynamic ocean tide model to obtain the ocean contribution in i. Other computations of ocean e¡ects have been proposed. Seiler & WÏnsch (1995, see also Brosche et al. 1989; Seiler 1989; Seiler 1991; WÏnsh & Seiler 1992) determined the short-term in£uences of ocean tides on UT1 from angular momentum conservation for a hydrodynamic model. Dickman (1993) computed the same quantity from a spherical harmonic tide theory (Dickman 1989, 1991). In view of the millimetre accuracy that is achieved by modern space geodetic techniques and the further improvements intended in the near future, a new improved model for earth rotation parameters is needed. Several models have been developed for precession and nutations (see Dehant & Defraigne 1997; Schastok 1997; Mathews et al. 1998); the present paper is concerned only with rotation rate variations. We compute the theoretical zonal response coe¤cient by using the same procedure as Wahr (1979) but for a more complete earth model. We investigate the e¡ect of modelling mantle inelasticity (the choice of reference period and frequency dependence). We also compute the e¡ect of a new initial state of the earth considering a convective mantle. This model contains departures from hydrostatic equilibrium; these are associated with (1) the density anomalies deduced from seismic velocity heterogeneities observed in the mantle by seismic tomography, (2) £ow-induced boundary deformations (internal as well as external), (3) the associated gravitational potential readjustment, and (4) a global earth dynamical £attening that is in agreement with the observed precession constant. We then add the e¡ect of ocean tides to the tidal variations of the lod thus obtained and compare the results with several kinds of observations. variable in generalized spherical harmonics. We then obtain a system of di¡erential equations of ¢rst order in d/dr. Due to the rotation and ellipticity of the Earth, these equations are coupled, but a system truncated at ¢rst order in the ellipticity can be solved. This is described in detail by Wahr (1979) and Dehant (1986). The system is then solved numerically from the earth's centre up to the surface, crossing each of the layers (inner core, outer core and di¡erent layers in the mantle) with suitable continuity conditions at each boundary. From this we obtain the solution for the displacements, stresses and potential readjustment at all depths in response to an external forcing potential. In particular, we obtain a toroidal displacement of degree 1 and order 0 as a response to a forcing potential of degree 2 and order 0. Its value at the surface can be interpreted in terms of the Earth's rotation velocity variation, and hence in terms of lod variations. Indeed, the toroidal displacement vector q01 can be written as an incremental rotation about the zª axis with respect to the mean rotation (see Wahr 1979): 2 L OD VA R IAT I O N S D U E TO T H E Z O NA L P OT E N T I A L *(lod)~ The response of the Earth to an external forcing is determined by the solution of the equation of motion, Poisson's equation for the Eulerian potential, and a stress^strain relation describing the rheology. These are vectorial equations that can be transformed into scalar equations by developing each q01 (u)~gs ei(utza) zªxr , (1) where gs is the angle of the rotation, u and a are the frequency and phase of the forcing, r is the vector giving the position of the point considered at the surface, and zª is the polar axis of the reference system, of which the origin is at the earth's instantaneous centre of mass and which is uniformly rotating with constant angular velocity )0 ~)0 zª. The norm of the vector q01 corresponds to the length of the circular arc at the Earth's surface, and can then be understood as a variation of UT1. Furthermore, the change in the Earth's rotation velocity is *)~ d(gs ei(utza) ) , dt (2) where d/dt is the time derivative. So, at the surface, we have *)~ 1 dq01 (u) r dt ~iugs ei(utza) : (3) Since d) produces a change in lod given by *(lod) *) ~{ , lod )0 (4) where lod~{2n/)0 , with )0 the mean rotation rate, we ¢nally obtain the lod variation in terms of the toroidal displacement: *(lod)~{ 2n iugs i(utza) e . )0 )0 (5) By using the relation between the lod variations and the universal time UT1, *(lod) d ~{ (UT1{UTC) , lod dt (6) and by considering the fact that we are looking for the response to a periodic exciting potential (of frequency u), we obtain 2in u(UT1{UTC) )0 (7) and (UT1{UTC)~{ ~ 1 *) iu )0 gs i(utza) e . )0 (8) ß 1999 RAS, GJI 139, 563^572 Length of day variations Note that in practice these expressions are computed as the response to a unit forcing, thus they correspond to transfer functions for *(lod) and (UT1{UTC). They must be multiplied by the amplitudes of a tide-generating potential (tgp) for each particular wave in order to obtain the ¢nal results in time units. Furthermore, it is common practice to express the transfer function for lod variations using the zonal response coe¤cient i, de¢ned by Agnew & Farrell (1978) as r *(lod) 1 5 a3 ~ i (9) V0 , lod 3 n GC 2 where V20 is the amplitude of the degree 2, order 0 tidal potential for a given frequency, G is the gravitational constant, C is the principal moment of inertia, and a is the earth's mean radius. In the case of an oceanless earth, the zonal response coe¤cient can be expressed as a function of the classical Love number k: i~ *Cm C k, *C Cm 1981) and the recent RATGP95 (Roosbeek 1996); our results for the hydrostatic case are presented in Table 1, where they are compared with Wahr's (1979) results. As can be seen from this table, the di¡erences in *lod are at most 2 per cent; it must be noted that this di¡erence is only due to the choice of initial earth model (PREM/1066A). We have veri¢ed that the choice of the tidal potential used could explain only a tenth of the di¡erence between the results (except for the wave Msf, for which it explains one half of the di¡erence). Also shown in Table 1 are the amplitudes derived from Yoder et al.'s (1981) results, using the value of k/C corresponding to an earth without ocean and with total decoupling between core and mantle (i.e. k/C~0.807). Their results are based on the Improved Lunar Ephemeris 1952^1959 (Eckert et al. 1954) and the value of the Love number k2 computed by Dahlen (1976) from model 1066A (Gilbert & Dziewonski 1975), which is the same as used by Wahr (1979). This is why their results are closer to Wahr's results than to ours. (10) where C and Cm are the principal moments of inertia of the earth and mantle respectively, and *C and *Cm are the corresponding increments due to the tidal deformations (see Nam & Dickman 1990 for more details). The results computed by Wahr (1979) correspond to model 1066A (Gilbert & Dziewonski 1975) with the hydrostatic equilibrium hypothesis. The tgp used was CTE (Cartwright & Tayler 1971; Cartwright & Edden 1973). We have made new computations with model PREM (Dziewonski & Anderson 3 NO N - H Y D RO STAT IC E QU I L I B R I U M MOD E L Based on several geophysical arguments, it is now accepted that the Earth is not in hydrostatic equilibrium. The ¢rst argument used is the existence of a non-hydrostatic geoid (Caputo 1965), which is also proof of the existence of mass heterogeneities inside the Earth. These mass anomalies are now observed by seismologists as lateral variations of seismic wave velocities and can be used to model the Earth with a Table 1. Amplitudes of the UT1 and lod variations induced by zonal tides (in ms). Tidal wave 18:6 yr Sa Ssa Msm Mm Msf Mf Mstm Mtm Msqm Mqm 565 UT1 lod UT1 lod UT1 lod UT1 lod UT1 lod UT1 lod UT1 lod UT1 lod UT1 lod UT1 lod UT1 lod 1 Period (days) Wahr (1979)1 This study2 hydrostatic Yoder et al: (1981)3 This study4 non-hydrostatic Effect of non-hydrostatic structure5 6798:37 {138:0 {0:127 1:31 0:0224 4:12 0:141 0:156 0:0307 0:706 0:160 0:063 0:0266 0:663 0:304 { { 0:085 0:0582 { { { { {136:0285 {0:1257 1:2923 0:0222 4:0690 0:1400 0:1539 0:0304 0:6970 0:1589 0:0620 0:0264 0:6540 0:3008 0:0166 0:0109 0:0837 0:0576 0:0104 0:0092 0:0083 0:0076 {138:8442 {0:1283 1:3186 0:0227 4:1421 0:1425 0:1566 0:0309 0:7092 0:1617 0:0630 0:0268 0:6659 0:3063 0:0169 0:0111 0:0852 0:0586 0:0106 0:0094 0:0085 0:0078 {136:0467 {0:1257 1:2925 0:0222 4:0698 0:1400 0:1539 0:0304 0:6971 0:1589 0:0620 0:0264 0:6541 0:3008 0:0166 0:0109 0:0837 0:0576 0:0104 0:0092 0:0083 0:0076 {0:0182 0:0 0:0002 0:0 0:0008 0:0 0:0001 0:0 0:0 0:0 0:0001 0:0 0:0 0:0 0:0 0:0 0:0 0:0 0:0 0:0 0:0 0:0 365:26 182:62 31:81 27:55 14:76 13:66 9:56 9:13 7:10 6:86 Corresponds to model 1066A with hydrostatic equilibrium and tidal potential CTE (Cartwright & Edden 1973). Corresponds to model PREM with hydrostatic equilibrium and tidal potential RATGP95 (Roosbeek 1996). 3 Corresponds to model 1066A with hydrostatic equilibrium and the improved lunar ephemeris 1952^1959 (Eckert et al. 1954). 4 Corresponds to model PREM with non-hydrostatic structure and RATGP95. 5 Di¡erence between columns (7) and (5). 2 ß 1999 RAS, GJI 139, 563^572 566 P. Defraigne and I. Smits non-hydrostatic equilibrium instead of the hydrostatic model in which all equidensity surfaces are elliptical. A second argument against hydrostatic equilibrium is the di¡erence of about 1 per cent between the dynamical £attening deduced from the observed precession constant and the theoretical value corresponding to hydrostatic equilibrium (Dehant & Capitaine 1997). Another argument concerns the free core nutation (FCN). This normal mode is due to the existence of a liquid rotating ellipsoidal core inside the solid rotating mantle, and to the fact that the rotation axes of these parts of the Earth can be misaligned. The FCN period is directly related to the core £attening. For a £attening corresponding to the hydrostatic equilibrium and to model PREM (Dziewonski & Anderson 1981), one ¢nds a period of 458 sidereal days in the inertial space [corresponding to a nearly diurnal period of {1/(1z1/458) days in a terrestrial frame]. The FCN induces a resonance in the diurnal tides and in nutations, so that it is possible to determine its period from the observed resonance. This leads to a value of 432 sidereal days in the inertial space (see e.g. Gwinn et al. 1986; Neuberg et al. 1987; Defraigne et al. 1994, 1995; Roosbeek et al. 1999). This disagreement between the theoretical and observed periods can be solved if we consider a core £attening about 5 per cent larger than the hydrostatic value (that is, an increase in the di¡erence between the core equatorial and polar radii of about 500 m). These arguments, among others, motivated us to recompute the Earth's response to the external gravitational potential due to the Moon, the Sun and the planets for a more realistic earth model accounting for the deviation from hydrostatic equilibrium observed. In two previous papers (Dehant & Defraigne 1997; Dehant et al. 1999) we have computed the response to the diurnal part of the potential (tesseral part of the tgp), leading to transfer functions for nutations (Dehant & Defraigne 1997) and tides (Dehant et al. 1999). In this latter paper, the Love numbers for the semi-diurnal and longperiod tides have also been computed. In the present paper, we apply the same procedure to obtain the response of the Earth to the long-period tides (zonal part of the tgp). This procedure consists of taking as a starting point the lateral variations in seismic velocity observed by seismic tomography. These velocity anomalies are then converted into density anomalies using a depth-dependent conversion factor (see Dehant & Defraigne 1997 for more details). The buoyancy-driven £ows associated with the mass anomalies are then deduced by modelling the steady-state mantle circulation, as presented in Defraigne et al. (1996), with a phase transition at 670 km depth. The viscosity pro¢le and the conversion factor between seismic velocity anomalies and density anomalies have been chosen inside acceptable bounds in order to have (1) a good correlation between the geoid and plate velocities deduced from the convection computation and the corresponding observed quantities, (2) global Earth dynamical £attening in agreement with the observed precession constant, and (3) a non-hydrostatic core £attening corresponding to what is necessary to have correspondence between the FCN period as deduced from geodetic and gravimetric observations and the theoretical FCN period. This constrained model was originally proposed by Defraigne (1997), but to avoid numerical instability in the computations of the transfer functions, Dehant & Defraigne (1997) slightly modi¢ed the conversion factor between seismic velocity anomalies and lateral density anomalies in order to remove the jumps in the £attenings of the equidensity surfaces at places other than the PREM discontinuities. The pro¢les thus obtained for the conversion factor and viscosity are shown in Dehant & Defraigne (1997) (Figs 4 and 5). We have applied the same parameter pro¢les for computing the lod transfer functions. From this convection model, we obtain the £owinduced boundary deformations, the £ow-induced potential readjustment and the £ow-induced mass readjustment in the inner core and outer core. We then take the degree 2, order 0 coe¤cients of all these quantities and of the lateral density heterogeneities deduced from the tomography model used as a starting point for the steady-state convection computation. Adding them to the £attening values of the hydrostatic model, we obtain new £attening values for each rheological property, for the potential and for the boundaries. This determines our initial earth model for computing the response to the external forcing. Note that in such a model the £attenings of the equidensity surfaces, the equipotential surfaces and the boundaries do not coincide, unlike in the case of the hydrostatic model. The results obtained for lod variations due to the zonal part of the tgp, for the non-hydrostatic model of the Earth, are given in Table 1. They correspond to a convolution between our computed transfer function and the model RATGP95 (Roosbeek 1996). When compared with the hydrostatic results (see Table 1), we see that the results are not a¡ected signi¢cantly by taking a non-hydrostatic initial earth model; the di¡erences between the results for the hydrostatic and nonhydrostatic models are less than 0.1 per cent of the total e¡ect, as seen in the last column of Table 1. The reason for this is that in a normal-mode expansion of the tidal solution (see Wahr 1979, 1981), the only important modes for the zonal tides are the axial spin modes (asm), corresponding to di¡erential toroidal motions between the di¡erent layers (inner core, outer core and mantle). These modes have a zero eigenfrequency (for a hydrostatic as well as a non-hydrostatic initial earth model), inducing a resonance in the long-period tides, and hence a frequency dependence of the solution, varying as 1/u. The asm resonance strengths are directly related to the change in C (the largest moment of inertia) induced by the Earth deformations in response to the tidal potential. However, while the inertia tensor is di¡erent for an earth in hydrostatic equilibrium and an earth in non-hydrostatic equilibrium, the variations in C due to the zonal tides are similar for both models. In the computations of the Earth's response to the zonal part of the tgp, the e¡ect of accounting for non-hydrostatic structure associated with mantle convection is thus negligible, as con¢rmed numerically by the results presented in Table 1. Note that this was not the case when computing the Earth response to the tesseral part of the tgp, that is, when computing diurnal tides (Dehant et al. 1999) and nutations (Dehant & Defraigne 1997). Indeed, in that case, two normalmode periods entering in the normal-mode expansion of the tidal solution are signi¢cantly a¡ected by the non-hydrostatic structure of the Earth. These are the Chandler wobble (CW) period (directly related to the global earth dynamical £attening) and the FCN period (directly related to the core £attening). Accounting for the non-hydrostatic structure induces changes in the eigenfrequencies of the CW and FCN and hence changes in the amplitudes of the tides and nutations with periods close to these resonance periods. Additionally, changes in the dynamical £attening and in the FCN frequency induce changes in the resonance strengths and hence changes in the tidal and nutation amplitudes. ß 1999 RAS, GJI 139, 563^572 Length of day variations 4 M A N T L E I N E L A ST ICI T Y The long-period tides, when corrected for oceanic and atmospheric e¡ects, can be used to constrain the frequency dependence of mantle inelasticity. Indeed, if the shear and bulk moduli and the corresponding attenuation factors Q depend on the frequency as ua (as proposed by Anderson & Minster 1979), where u is the frequency and a is the inelastic parameter, the response of the earth increases for decreasing frequencies (or increasing periods). In this study, we propose to apply the same frequency dependence for computing the mantle inelasticity e¡ects on lod variations induced by the long-period tides. Our results are thus based on Anderson & Minster's (1979) model, described by a na na u Q(u)~ cot z Q(u0 ){ cot , (11) 2 2 u0 p Q(u) Q2 (u0 )z1 u a p(u)~p(u0 ) , (12) u0 Q2 (u)z1 where p(u) is written for the bulk modulus k(u) or for the shear modulus k(u). The parameter a of this frequency dependence has been evaluated from several kinds of observations at long periods. From the observed Chandler wobble, Smith & Dahlen (1981) found that reasonable values are between 0.1 and 0.2. From the observed value of the zonal response coe¤cient i, Merriam (1984) provided an upper bound for a of 0.3. Wahr & Bergen (1986) used the results of Smith & Dahlen (1981) to compute the e¡ect of the inelasticity e¡ect on lod variations for longperiod tides. The models of Smith & Dahlen (1981) correspond to (1) a~0:15 for the Q values of the seismic model QMU of Sailor & Dziewonski (1978) and (2) a~0:09 for the model b of Sipkin & Jordan (1980), with reference frequencies equal to 200 s for model QMU and 30 s for model b. Because of the frequency dependence of the parameters, the most e¡ective wave for constraining the a value is the 18.6 yr wave. Recently, Eanes (1995; see also Eanes & Bettadpur 1997) has derived mantle inelasticity properties from satellite observations of the Love number k2 for the 18.6 yr tide; using the Anderson & Minster (1979) model with a reference period of 1 s, Eanes deduced a value of a~0:14. In this study, we choose a value of a~0.15 as in Dehant & Defraigne (1997). This value has been deduced from a comparison between the Love number k2 computed for the 18.6 yr tide, using the Anderson & Minster (1979) model with a reference period of 200 s, and the observed value given by Eanes & Bettadpur (1997). In order to investigate the importance of the parameters (other than a) in the inelastic model, we have tested di¡erent hypotheses for the starting model for the Q pro¢le. First we used the Q values from PREM at 1 s; second we used the same values but with 200 s as the reference period; and third we used the more recent Q values given by Widmer et al. (1991) at 300 s. We propose, as in Dehant & Defraigne (1997), di¡erent schemes of inelastic parametrizations; these are illustrated in Fig. 1. Model A uses the Q values of PREM with a reference period of 200 s to obtain the bulk and shear moduli at the tidal period from the values given by PREM at a period of 1 s. Model B uses the same Q values to obtain the shear and bulk moduli at a period of 200 s but uses the Q values of Widmer et al. (1991) to obtain the bulk and shear moduli at the tidal period from the values of these parameters at 200 s. Finally, model C is the same as model B but assuming that the PREM Q values correspond to a reference period of 1 rather than 200 s. In the three models proposed, we use the logarithmic frequency dependence between periods of 1 and 200 s proposed by Dziewonski & Anderson (1981) for model PREM; this corresponds to the following relations: 2 u , (13) Q(u)~Q(u0 ){ ln n u0 p Q(u) Q2 (u0 )z1 p(u)~p(u0 ) , (14) Q2 (u)z1 where p(u) is written for k(u) or k(u). This logarithmic law is represented by dashed lines in Fig. 1. The solid lines correspond to the frequency dependence as ua (eqs 11 and 12), used for going from 200 s up to the tidal period. Table 2 shows the numerical values obtained for models A, B and C with a~0:15, and for model B with a~0:10. It can be seen that the e¡ect of taking Widmer's Q values rather Figure 1. Scheme of the way in which the inelasticity models A, B and C are constructed. The solid lines show the frequency dependence as ua , and the dashed lines show the logarithmic law. QW are the Q values of Widmer et al. (1991). Table 2. E¡ects of mantle inelasticity on UT1 variations (ms). Tidal wave Period (days) 18:6 yr Sa Ssa Mm Mf ß 1999 RAS, GJI 139, 563^572 6798 365:25 182:62 27:55 13:66 567 Model A a~0:15 This study Model B Model B a~0:15 a~0:10 {10:553 0:0647 0:1828 0:0234 0:0197 {9:737 0:0599 0:1692 0:0218 0:0184 Model C a~0:15 {6:571 {9:724 0:0452 0:0598 0:1317 0:1688 0:0181 0:0217 0:0156 0:0183 Wahr & Bergen (1986) Model QMU Model b a~0:15 a~0:09 { 0:0530 0:1470 0:0176 0:0143 { 0:0651 0:1875 0:0249 0:0211 568 P. Defraigne and I. Smits than PREM's Q values in the a-law (comparison between the results of models A and B) is more important than the choice of the reference period for QPREM (comparison between the results of models B and C). However, the numerical value of a has the most important impact on the inelastic contribution, as seen from a comparison between the results for model B with a~0.15 and a~0.10. Also shown in Table 2 are the corresponding results obtained by Wahr & Bergen (1986) for models QMU and b. The reasons why they obtain such di¡erent results for the same value of a (a~0.15, model QMU) are due to the facts that ¢rst these authors use di¡erent sets of Q values, and second they use only the a-law (as illustrated in Fig. 2), while we use the logarithmic law in the seismic frequency band, i.e. between 1 and 200 s, as proposed by Dziewonski & Anderson (1981) for PREM. As shown in Table 2, these choices induce signi¢cant e¡ects on the numerical results of UT1 variations. Our preferred model is model B because PREM is computed from a set of normal-mode and body-wave observations in the seismic frequency band for which the mean period is 200 s, and because it uses the more recent Q values of Widmer et al. (1991). In that frame, from a comparison with the observed 18.6 yr tidal mass redistribution potential (deduced from Eanes & Bettadpur's 1997 value), our preferred value for a is 0.15. 5 C OMPA R IS O N W IT H GE OD ET IC O B S E RVAT I O N S We have taken from the literature three di¡erent series of observed lod variations for the main long-period tidal waves. All of these values correspond to least-squares estimations of geodetic observations, after removing the atmospheric contributions. First, we use the observed i from Chao et al. (1995), who determined their results from the `Space92' universal timeseries (Gross 1993), which results from a Kalman ¢lter combination of all available space geodetic measurements of Earth orientation. The AAM series used to extract the atmospheric excitation were from the ECMWF (European Centre for Medium-Range Weather Forecasts) for the ¢rst part of the series and from the JMA (Japan Meteorological Agency) for the second part. A second set of observed i was derived from the results of Robertson et al. (1994). These authors deduced their results Figure 2. Scheme of the way in which the inelasticity models b and QMU are constructed by Wahr & Bergen (1986). The solid lines show the frequency dependence as ua . from VLBI data and provided values for k/C as proposed by Yoder et al. (1981), where k is the tidal e¡ective (that is, concerning only the parts of the Earth involved in UT1 variationsöthe mantle and crust) gravitational Love number and C is the dimensionless polar moment of inertia (i.e. corresponding to C/Ma2, where M is the Earth's mass and a is the mean equatorial radius). The k/C estimations are determined by a least-squares ¢t on the VLBI observations, ¢rst corrected for AAM e¡ects using AAM series from the NMC (US National Meteorological Center), with and without correction associated with the inverted barometer e¡ect, from the JMA, with and without correction associated with the inverted barometer e¡ect, and from the ECMWF. We then multiplied the k/C values by 0.333 (the dimensionless moment of inertia of the Earth) to obtain the corresponding i value. Finally, we also compared our results with the values computed by Dickman & Nam (1995) from IRIS (International Radio Interferometric Surveying) UT1 data, ¢rst corrected for atmospheric e¡ects using the NMC AAM time-series, with and without the e¡ect of the inverted barometer. These three sets of data are given in Table 3. As can be seen in this table, the di¡erent AAM series used for correcting the atmospheric contribution lead to di¡erences of up to 10 per cent in the values computed for i; all the data sets thus provide us with the possible range of i values derived from observations. Before comparing these observed values with our theoretical computations, ocean corrections must be accounted for (the ocean contribution is at the 10 per cent level). For this, we use ¢rst the corrections computed by Dickman (1993) from a spherical harmonic tide theory (Dickman 1991) for a dynamic ocean, and second the ocean corrections given by Seiler & WÏnsch (1995) from a numerical dynamic ocean model. These corrections are presented in Table 4. It can be seen that the results of Dickman (1993) lead to corrections about 10^30 per cent larger than the results of Seiler & WÏnsch (1995). For the waves for which no correction was provided, we have interpolated the existing values to obtain an estimate of the e¡ect at the corresponding periods (see values in italics in Table 4). Finally, we give in Table 5 the theoretical values computed for our preferred earth model, that is, the model initially in non-hydrostatic equilibrium, with mantle inelasticity parameters corresponding to model B with a~0:15 and corrected for ocean e¡ects; both sets of ocean corrections discussed above were used. In Table 5 we also compare our results with the two models provided by the IERS conventions (McCarthy 1996), which are the theoretical UT1 variations from Yoder et al. (1981), corresponding to an elastic earth with decoupling between the core and the mantle. Model UT1R incorporates the ocean corrections of Yoder et al. (1981), while model UT1S incorporates the ocean corrections of Dickman (1993), except for the waves Mqm, Mstm and Mstm, for which Yoder et al.'s (1981) values are used because these waves are not given by Dickman (1993). Note that Yoder et al.'s (1981) corrections consist of one constant value added to the Love number k, and corresponding to the equilibrium ocean tide contribution given by Agnew & Farrell (1978) for model 1066A. In Fig. 3, we present a comparison between these theoretical models. We added our numerical results obtained for model B with a~0:10 in order to allow a better visualization of the in£uence of this important parameter in the modelling of inelasticity. Note that, for our theoretical results, we have joined the di¡erent points of Table 5 in order to show the frequency ß 1999 RAS, GJI 139, 563^572 Length of day variations 569 Table 3. Observed values for i (in-phase part, with standard deviation) corrected for atmospheric e¡ects. Tidal wave Period (days) Ssa 182:62 Msm 31:81 Mm 27:55 Msf 14:76 Mf 13:66 Mstm 9:56 Mtm 9:13 Msqm 7:10 Mqm 6:86 Chao et al: (1995) 0:3323 0:0450 0:3081 0:0067 0:3359 0:0027 0:3103 0:0018 0:3249 0:0306 0:3001 0:0062 0:3024 0:0218 0:3075 0:0280 NMC Robertson et al: (1994) NMCzIB JMA JMAzIB ECMWF 0:3522 0:0195 0:3405 0:0145 0:3149 0:0002 0:3137 0:0002 0:3077 0:0002 0:3075 0:0002 0:3052 0:0002 0:3144 0:0235 0:3134 0:0203 0:3114 0:0002 0:3132 0:0002 0:3113 0:0002 0:3129 0:0002 0:3138 0:0002 0:3021 0:0098 0:3021 0:0098 0:3271 0:0014 0:3149 0:0014 0:3559 0:0014 0:3414 0:0014 0:3565 0:0014 0:2583 0:0516 0:2507 0:0565 Table 4. Ocean corrections for i (in-phase). Tidal wave Period (days) Dickman (1993) Seiler & W unsch (1995) 18:6 yr Sa Ssa Msm Mm Msf Mf Mstm Mtm Msqm Mqm 6798:37 365:26 182:62 31:81 27:55 14:76 13:66 9:56 9:13 7:10 6:86 0:0489 0:0490 0:0489 0:0476 0:0475 0:0445 0:0439 0:0401 0:0397 0:0378 0:0376 0:0440 0:0440 0:0438 0:0389 0:0397 0:0339 0:0346 0:0316 0:0283 0:0255 0:0251 Values in italics correspond to interpolations as a function of the frequency. dependence; the curves are not perfectly regular because of the ocean corrections (which do not have a smoothed frequency dependence). In Fig. 3, we have also plotted the values of i corresponding to the IERS96 models UT1R and UT1S. The strong singularities in the corresponding curves (for the waves Mqm and Mstm) are due to the ocean corrections, which are identical at these frequencies, as explained above. In our models, we interpolated the model of Dickman (1993) as a function of the frequency to obtain the ocean corrections for these wavesöthis is why we obtain a frequency dependence of i that is much smoother than in models UT1R and UT1S. The di¡erences between models UT1R and UT1S and our results are due ¢rst to the fact that the IERS models are based on an elastic mantle, while we account for mantle inelasticity, and second to the di¡erent ocean corrections. In Fig. 3, it is clear that the di¡erence between the results for a~0:10 and a~0:15 is much smaller than the di¡erence between the results obtained with the two ocean correction models, and that the di¡erences between the ocean correction models are even larger than the inelastic contribution that is absent from models UT1R and UT1S. ß 1999 RAS, GJI 139, 563^572 Dickman & Nam (1995) without IB with IB Fig. 4 presents a comparison between our new theoretical model and some observations corrected for atmospheric e¡ects and shown with 1p error bars. It is clear that the precision of the observations does not allow us to infer any information about the parameter a. It is noted that the e¡ect of mantle inelasticity is mostly visible at the frequency of the 18.6 yr tide, but the tidal contribution at that period is not easily observed because of the poor knowledge about the relative contributions of tidal deformation and core angular momentum to the Earth's rotation. Moreover, as seen in Fig. 4, the precision of the observations is not su¤cient to discriminate between the two sets of ocean corrections used in this study. 6 CO NC LUSI O N S By solving numerically the equation of motion, the potential equation and the stress^strain relation inside the Earth, we have computed new transfer functions for the lod variations induced by the zonal part of the tide-generating potential. In this computation, we have investigated the e¡ect of the earth modelling for the initial state (before tidal deformation). On the one hand, we have shown that accounting for nonhydrostatic structure associated with steady-state mantle convection (including lateral heterogeneities as well as deformed boundaries) has an e¡ect of less than 0.1 per cent on the results. On the other hand, we have estimated the in£uence of mantle inelasticity and shown the importance of its modelling (the choice of a model for Q, the choice of the reference period, the choice of the frequency dependence inside the seismic frequency band and the choice of the parameter a). We have shown that the two most important parameters are (1) the frequency dependence law in the seismic frequency band, and (2) the value of a for the frequency dependence outside this frequency band. In this study, we chose the logarithmic law inside the seismic frequency band and a~0:15 for the frequency dependence as ua (see eqs 11 and 12) outside the seismic frequency band. Finally, we have proposed a re¢ned theoretical model for the lod variations induced by the zonal part of the 570 P. Defraigne and I. Smits Table 5. Comparison among theoretical results corrected for ocean e¡ects. This study (D)1 a~0:15 Tidal wave Period (days) 18:6 yr Sa Ssa Msm Mm Msf Mf Mstm Mtm Msqm Mqm 6798:37 365:26 182:62 31:81 27:55 14:76 13:66 9:56 9:13 7:10 6:86 0:3296 0:3242 0:3229 0:3191 0:3188 0:3150 0:3143 0:3102 0:3097 0:3076 0:3073 18:6 yr Sa Ssa Msm Mm Msf Mf Mstm Mtm Msqm Mqm 6798:37 365:26 182:62 31:81 27:55 14:76 13:66 9:56 9:13 7:10 6:86 {171:0527 1:5926 4:9947 0:1867 0:8447 0:0742 0:7816 0:0196 0:0986 0:0121 0:0097 This study (SW)2 a~0:15 UT1R UT1S values for i 0:3247 0:3192 0:3178 0:3104 0:3110 0:3044 0:3050 0:3017 0:2983 0:2953 0:2948 { { { 0:3111 0:3117 0:3099 0:3121 0:3162 0:3111 0:3039 0:3161 0:3156 0:3165 0:3157 0:3145 0:3144 0:3056 0:3109 0:3162 0:3079 0:3039 0:3161 corresponding amplitudes for UT1 in ms 1 2 {168:5101 1:5681 4:9158 0:1816 0:8240 0:0717 0:7584 0:0191 0:0949 0:0117 0:0093 { { { 0:182 0:826 0:073 0:776 0:020 0:099 0:012 0:010 {163:768 1:555 4:884 0:184 0:833 0:072 0:773 0:020 0:098 0:012 0:010 Ocean correction of Dickman (1993). Ocean correction of Seiler & WÏnsch (1995). tgp, which accounts for the Earth's non-hydrostatic equilibrium, for mantle inelasticity and for ocean e¡ects. This model is an improvement with respect to the IERS96 models UT1R and UT1S because these two models do not include the mantle inelasticity e¡ects, and both use (for all or for particular frequencies) a set of ocean corrections from Yoder et al. (1981), which are based on a simple estimation of the ocean loading Love number that is constant with frequency. In our model, we Figure 3. Comparison between our theoretical model for i corrected for ocean e¡ects and models UT1R and UT1S of the IERS96 conventions. ß 1999 RAS, GJI 139, 563^572 Length of day variations 571 Figure 4. Comparison between our theoretical model for i corrected for ocean e¡ects and the values of i deduced from geodetic observations. used the ocean corrections of Seiler & WÏnsch (1995) and Dickman (1993) instead, which are based on dynamic ocean models. Our major conclusion is that the precision of the geodetic observations of UT1 and the precision of the oceanic and atmospheric corrections are not yet su¤cient to constrain the parameter a involved in the inelasticity model from a comparison of our theoretical results with these observations. This is why we have used the value of a~0:15 that has been determined from a comparison between the Love number k2 computed for the 18.6 yr tide and the corresponding observations of the variation in the gravitational potential deduced from satellite observation (Eanes & Bettadpur 1997). As a last remark, we note that the results presented here concern only the in-phase part of the UT1 variations associated with the zonal tides. A non-negligible out-of-phase part is observed (see the references of the observational results given in Section 4); it is due to the dissipation associated with mantle inelasticity (a numerical estimation is provided by Wahr & Bergen 1986) and to the ocean e¡ects. Another part probably corresponds to the atmospheric e¡ect, which is not perfectly corrected for because of the low precision of the atmospheric models. The non-hydrostatic structure of the Earth, which we have included in this study, has no e¡ect on the out-of-phase part. AC K NOW L E D GM E N T S We would like to thank Veronique Dehant for the scienti¢c discussions and for her careful reading of the paper. The numerical results of this paper have been obtained with a program derived from John Wahr's program for an ellipsoidal earth. We also thank Harald Schuh and an anonymous reviewer for their very constructive review of the paper. ß 1999 RAS, GJI 139, 563^572 R E F ER E NC E S Agnew, D.C. & Farrell, W.E., 1978. Self-consistent equilibrium ocean tides, Geophys. J. R. astr. Soc., 55, 171^181. Anderson, D.L. & Minster, J.B., 1979. The frequency dependence of Q in the Earth and implications for mantle rheology and Chandler wobble, Geophys. J. R. astr. Soc., 58, 431^440. Brosche, P., Seiler, U., SÏndermann, J. & WÏnsch, J., 1989. Periodic changes in Earth's rotation due to oceanic tides, Astron. Astrophys., 220, 318^320. Caputo, 1965. The minimum strength of the Earth, J. geophys. Res., 70, 955^963. Cartwright, D.E. & Edden, A.C., 1973. Corrected tables of tidal harmonics, Geophys. J. R. astr. Soc., 33, 253^264. Cartwright, D.E. & Tayler, R.J., 1971. New computations of the tide generating potential, Geophys. J. R. astr. Soc., 23, 45^74. Chao, B.F., Merriam, J.B. & Tamura, Y., 1995. Geophysical analysis of zonal tidal signals in length of day, Geophys. J. Int., 122, 765^775. Dahlen, F.A., 1976. The passive in£uence of the oceans upon the rotation of the Earth, Geophys. J. R. astr. Soc., 46, 363^406. Defraigne, P., 1997. Geophysical model of the Earth dynamical £attening in agreement with the precession constant, Geophys. J. Int., 130, 47^56. Defraigne, P., Dehant, V. & Hinderer, J., 1994. Stacking gravity tide measurements and nutation observations in order to determine the complex eigenfrequency of the nearly diurnal free wobble, J. geophys. Res., 99, 9203^9213. Defraigne, P., Dehant, V. & Hinderer, J., 1995. Correction to `Stacking gravity tide measurements and nutation observations in order to determine the complex eigenfrequency of the nearly diurnal free wobble', J. geophys. Res., 100, 2041^2042. Defraigne, P., Dehant, V. & Wahr, J., 1996. Internal loading of a homogeneous compressible Earth with phase boundaries, Geophys. J. Int., 125, 173^192. Dehant, V., 1986. Intëgration des ëquations aux dëformations d'une Terre elliptique, inëlastique, en rotation uniforme et a© noyau liquide, PhD thesis, Universitë Catholique de Louvain, Louvain-la-Neuve, Belgium (in French). 572 P. Defraigne and I. Smits Dehant, V. & Capitaine, N., 1997. On the luni-solar precession, the tilt-over-mode, and the Oppolzer terms, Celest. Mech., 65, 439^458. Dehant, V. & Defraigne, P., 1997. New transfer functions for nutations of a non-rigid Earth, J. geophys. Res., 102, 27 659^27 687. Dehant, V., Defraigne, P. & Wahr, J.M., 1999. Tides for a convective Earth, J. geophys. Res., 104, 1035^1058. Dickman, S.R., 1989. A complete spherical harmonic approach to luni-solar tides, Geophys. J., 99, 457^468. Dickman, S.R., 1991. Ocean tides for satellite geodesy, Mar. Geod., 14, 21^56. Dickman, S.R., 1993. Dynamic ocean tide e¡ects on Earth's rotation, Geophys. J. Int., 112, 448^470. Dickman, S.R. & Nam, Y.S., 1995. Revised predictions of lon-period ocean tidal e¡ects on Earth's rotation rate, J. geophys. Res., 100, 8233^8243. Dziewonski, A.D. & Anderson, D.L., 1981. Preliminary reference earth model, Phys. Earth planet. Inter., 25, 297^356. Eanes, R.J., 1995. A study of temporal variations in Earth's gravitational ¢eld using Lageos-1 laser range observations, PhD thesis, CSR-95-8, Center for Space Research, University of Texas, Austin. Eanes, R.J. & Bettadpur, S.V., 1997. Temporal variability of Earth's gravitational ¢eld from satellite laser ranging, in Global Gravity Field and its Variations, Proc. Symp. 116 G3, IAG Symp. Series, eds Rapp, R.H., Cozenave, A.A. & Nerem, R.S., Springer. Eckert, W.J., Jones, R. & Clark, H.K., 1954. Construction of the lunar ephemeris, in Improved Lunar Ephemeris 1952-1959, pp. 283^363, US Govt Printing O¤ce, Washington, DC. Gilbert, F. & Dziewonski, A.M., 1975. An application of normal mode theory to the retrieval of structure parameters and source mechanisms for seismic spectra, Phil. Trans. R. Soc. Lond., A278, 187^269. Gross, R.S., 1993. A combination of Earth orientation data: Space 92, JPL Tech. Memo., Jet Propulsion Laboratory, Pasadena, CA. Gwinn, C.R., Herring, T.A. & Shapiro, I.I., 1986. Geodesy by radio interferometry: studies of the forced nutations of the Earth, 2. Interpretation, J. geophys. Res., 91(B5), 4755^4765. Hefty, J. & Capitaine, N., 1990. The fortnightly and monthly zonal tides in the Earth's rotation from 1962 to 1988, Geophys. J. Int., 103, 219^231. Mathews, P.M., Bu¡ett, B.A. & Herring, T.A., 1998. The magnetic coupling contribution to nutation, in Proc. Journës Syste©mes de Rëfërence 1998, Paris, France, in press. McCarthy, D.D., 1996. IERS Conventions, IERS Technical Note, 21, ed. McCarthy, D.D., Observatoire de Paris. McCarthy, D.D. & Luzum, B.J., 1993. An analysis of tidal variations in the length of day, Geophys. J. Int., 114, 341^346. Merriam, J.B., 1980. Zonal tides and changes in the length of day, Geophys. J. R. astr. Soc., 62, 551^556. Merriam, J.B., 1984. Tidal terms in universal time: e¡ects of zonal winds and mantle Q, J. geophys. Res., 89 (B12), 10 109^10 114. Merriam, J.B., 1985. Lageos and UT measurements of long-period Earth tides and mantle Q, J. geophys. Res., 90, 9423^9430. Nam, Y.S. & Dickman, S.R., 1990. E¡ect of dynamic long-period ocean tides on changes in Earth's rotation rate, J. geophys. Res., 95, 6751^6757. Neuberg, J., Hinderer, J. & ZÏrn, W., 1987. Stacking gravity tide observations in central Europe for the retrieval of the complex eigenfrequency of the nearly diurnal free wobble, Geophys. J. R. astr. Soc., 91, 853^868. Robertson, D.S., Ray, J.R. & Carter, W.E., 1994. Tidal variations in UT1 observed with very long baseline interferometry, J. geophys. Res., 99, 621^636. Roosbeek, F., 1996. A tide generating potential precise to the nanogal level, Geophys. J. Int., 126, 197^204. Roosbeek, F., Defraigne, P., Feissel, M. & Dehant, V., 1999. The free core nutation period stays between 431 and 434 sidereal days, Geophys. Res. Lett., 26, 131^134. Sailor, R.V. & Dziewonski, A.M., 1978. Measurements and interpretation of normal mode attenuation, Geophys. J. R. astr. Soc., 53, 559^581. Schastok, J., 1997. A new nutation series for a more realistic model Earth, Geophys. J. Int., 130, 137^150. Schastok, J., So¡el, M. & Ruder, H., 1994. A contribution to the study of fortnightly and monthly zonal tides in UT1, Astron. Astrophys., 283, 650^654. Seiler, U., 1989. Variations of the angular momentum budget for tides of the present ocean, in Earth's Rotation from Eons to Days, pp. 81^94, eds Brosche, P. & SÏndermann, J., Springer-Verlag, New York. Seiler, U., 1991. Periodic changes of the angular momentum budget due to tides of the world ocean, J. geophys. Res., 96, 10 287^10 300. Seiler, U. & WÏnsch, J., 1995. A re¢ned model for the in£uence of ocean tides on UT1 and polar motion, Astr. Nachr., 316, 419. Sipkin, S.A. & Jordan, T.H., 1980. Frequency dependence of QScS , Bull. seism. Soc. Am., 70, 1071^1102. Smith, M.L. & Dahlen, F.A., 1981. The period and Q of the Chandler wobble, Geophys. J. R. astr. Soc., 64, 223^281. Wahr, J.M., 1979. The tidal motions of a rotating, elliptical, elastic and oceanless Earth, PhD thesis, University of Colorado, Boulder. Wahr, J.M., 1981. A normal mode expansion for the forced response of rotating Earth, Geophys. J. R. astr. Soc., 64, 651^675. Wahr, J.M. & Bergen, Z., 1986. The e¡ects of mantle anelasticity on nutations, Earth tides, and tidal variations of rotation rate, Geophys. J. R. astr. Soc., 87, 633^668. Wahr, J.M., Sasao, T. & Smith, M.L., 1981. E¡ect of the £uid core on changes in the length of day due to long period tides, Geophys. J. R. astr. Soc., 64, 635^650. Widmer, R., Masters, G. & Gilbert, F., 1991. Spherically symmetric attenuation within the Earth from normal mode data, Geophys. J. Int., 104, 541^553. WÏnsch, J. & Seiler, U., 1992. Theoretical amplitudes and phases of the periodic polar motion terms caused by ocean tides, Astron. Astrophys., 226, 581^587. Yoder, C.F., Williams, J.G. & Parke, M., 1981. Tidal variations of Earth rotation, J. geophys. Res., 86, 881^891. ß 1999 RAS, GJI 139, 563^572
© Copyright 2026 Paperzz