17
Geophys. J. Int. (2007) 171, 780–796 doi: 10.1111/j.1365-246X.2007.03553.x GJI Seismology On reduction of long-period horizontal seismic noise using local barometric pressure W. Z¨ urn, 1 J. Exß, 2 H. Steffen, 3 C. Kroner, 4 T. Jahr 4 and M. Westerhaus 5 1 Black Forest Observatory (Schiltach), Universities Karlsruhe/Stuttgart, Heubach 206, D-77709 Wolfach, Germany. E-mail: [email protected] 2 Thinking Objects GmbH, Stuttgart, Germany 3 Institut f ¨ ur Erdmessung, University of Hannover, Schneiderberg 50, D-30167 Hannover, Germany 4 Institute of Geosciences, University of Jena, Burgweg 11, D-07749 Jena, Germany 5 Geodetic Institute, University of Karlsruhe, Englerstr. 7, D-76128 Karlsruhe, Germany Accepted 2007 July 23. Received 2007 July 11; in original form 2006 August 31 SUMMARY Tilt from atmospheric loading has long been known to be the major source of long-period horizontal seismic noise. We try to quantify these effects for seismic data from the Black Forest Observatory (BFO), which is known to be a very quiet station. Experimental transfer functions between local barometric pressure and horizontal seismic noise are estimated for two long time-series by standard methods. Two simple analytical physical models are developed: the local deformation model (LDM) and the acoustic-gravity wave model (TWM). Subsequently these models, with only two free parameters are fit using least squares to the observed seismic noise for time-series of widely differing lengths. The results are variable, sometimes rather dramatic variance reductions are obtained and sometimes the reduction is hardly significant. The method produces the best results when barometrically induced noise is high. The resulting admittances for the LDM are compared to finite element calculations. Since the methods are simple and can result in conspicuous reductions in noise we provide one more reason for installing barometers at even the best broad-band seismic stations. Key words: barometric pressure, horizontal seismic noise, modelling, tilt. 1 INTRODUCTION It is a long-known fact that horizontal seismic noise below 50 mHz is larger than the vertical noise by factors between 5 and 15 (e.g. M¨ uller & Z ¨ urn 1983, Fig. 6). This is confirmed in a recent study by Berger et al. (2004) of the noise at many stations of the Global Seismic Network (GSN). It is also generally accepted that most of this excessive horizontal noise is caused by tilts generated by atmospheric pressure changes and wind (e.g. Webb 2002; Wielandt 2002). Tilts were also invoked as the responsible mechanism by Tsai et al. (2004) who studied the propagation of atmospheric morning glory waves in the Los Angeles basin with speeds between 5 and 25 ms 1 , frequencies near 1 mHz, and amplitudes between 0.8 and 1.3 hPa by using records from barometers and seismometers. These authors note also that the amplitudes on the horizontal seismometers were clearly larger. At sites near coasts loading from the ocean waves causes a tem- porally varying deformation field (with tilts) which is sensed by seismographs. A fine example of this sensitivity was demonstrated for the load tilts generated at coastal sites in the Indian ocean by the tsunamis radiated by the N. Sumatra—Andaman Islands quake on 2004 December 26 as studied by Yuan et al. (2005). These are near-field effects (also invoked in a study of vertical seismic noise by Agnew & Berger 1978) in contrast to the well-known marine microseisms, that is, propagating surface waves excited by waves in the oceans. However, in this paper, we deal with the long-period horizontal noise at the best stations, for well-shielded sensors and far away from the coasts. We note here in passing that the atmosphere also is able to generate propagating surface waves. Manifestations of this could be the background free oscillations first reported by Nawa et al. (1998) and the surface waves observed during the parox- ysmal eruptions of El Chich ´ on in 1982 and Mount Pinatubo in 1991 (e.g. Z ¨ urn & Widmer 1996). Barometric pressure changes in the immediate environment of the seismic instruments can lead to noise by mechanisms like buoy- ancy (vertical component), deformation and tilt of the instrument housing. Air convection has been identified as a major source of noise a long time ago. Another source of noise on seismographs with helical and leaf springs has recently been identified and stud- ied by Forbriger (2007): variations of Earth’s magnetic field. We do not consider these effects here because we are interested in the noise which cannot be avoided by clever construction, shielding and installation of the sensors. Tilt contributions to the noise in the free mode band associ- ated with changes of barometric pressure were already studied by Große-Brauckmann (1979) in the context of borehole installations 780 C 2007 The Authors Journal compilation C 2007 RAS

On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

  • Upload
    others

  • View
    11

  • Download
    0

Embed Size (px)

Citation preview

Page 1: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

Geophys. J. Int. (2007) 171, 780–796 doi: 10.1111/j.1365-246X.2007.03553.xG

JISei

smol

ogy

On reduction of long-period horizontal seismic noise using localbarometric pressure

W. Zurn,1 J. Exß,2 H. Steffen,3 C. Kroner,4 T. Jahr4 and M. Westerhaus5

1Black Forest Observatory (Schiltach), Universities Karlsruhe/Stuttgart, Heubach 206, D-77709 Wolfach, Germany.E-mail: [email protected] Objects GmbH, Stuttgart, Germany3Institut fur Erdmessung, University of Hannover, Schneiderberg 50, D-30167 Hannover, Germany4Institute of Geosciences, University of Jena, Burgweg 11, D-07749 Jena, Germany5Geodetic Institute, University of Karlsruhe, Englerstr. 7, D-76128 Karlsruhe, Germany

Accepted 2007 July 23. Received 2007 July 11; in original form 2006 August 31

S U M M A R YTilt from atmospheric loading has long been known to be the major source of long-periodhorizontal seismic noise. We try to quantify these effects for seismic data from the BlackForest Observatory (BFO), which is known to be a very quiet station. Experimental transferfunctions between local barometric pressure and horizontal seismic noise are estimated for twolong time-series by standard methods. Two simple analytical physical models are developed: thelocal deformation model (LDM) and the acoustic-gravity wave model (TWM). Subsequentlythese models, with only two free parameters are fit using least squares to the observed seismicnoise for time-series of widely differing lengths. The results are variable, sometimes ratherdramatic variance reductions are obtained and sometimes the reduction is hardly significant.The method produces the best results when barometrically induced noise is high. The resultingadmittances for the LDM are compared to finite element calculations. Since the methods aresimple and can result in conspicuous reductions in noise we provide one more reason forinstalling barometers at even the best broad-band seismic stations.

Key words: barometric pressure, horizontal seismic noise, modelling, tilt.

1 I N T RO D U C T I O N

It is a long-known fact that horizontal seismic noise below

∼50 mHz is larger than the vertical noise by factors between 5 and

15 (e.g. Muller & Zurn 1983, Fig. 6). This is confirmed in a recent

study by Berger et al. (2004) of the noise at many stations of the

Global Seismic Network (GSN). It is also generally accepted that

most of this excessive horizontal noise is caused by tilts generated by

atmospheric pressure changes and wind (e.g. Webb 2002; Wielandt

2002). Tilts were also invoked as the responsible mechanism by Tsai

et al. (2004) who studied the propagation of atmospheric morning

glory waves in the Los Angeles basin with speeds between 5 and

25 ms−1, frequencies near 1 mHz, and amplitudes between 0.8 and

1.3 hPa by using records from barometers and seismometers. These

authors note also that the amplitudes on the horizontal seismometers

were clearly larger.

At sites near coasts loading from the ocean waves causes a tem-

porally varying deformation field (with tilts) which is sensed by

seismographs. A fine example of this sensitivity was demonstrated

for the load tilts generated at coastal sites in the Indian ocean by

the tsunamis radiated by the N. Sumatra—Andaman Islands quake

on 2004 December 26 as studied by Yuan et al. (2005). These are

near-field effects (also invoked in a study of vertical seismic noise

by Agnew & Berger 1978) in contrast to the well-known marine

microseisms, that is, propagating surface waves excited by waves

in the oceans. However, in this paper, we deal with the long-period

horizontal noise at the best stations, for well-shielded sensors and far

away from the coasts. We note here in passing that the atmosphere

also is able to generate propagating surface waves. Manifestations

of this could be the background free oscillations first reported by

Nawa et al. (1998) and the surface waves observed during the parox-

ysmal eruptions of El Chichon in 1982 and Mount Pinatubo in 1991

(e.g. Zurn & Widmer 1996).

Barometric pressure changes in the immediate environment of

the seismic instruments can lead to noise by mechanisms like buoy-

ancy (vertical component), deformation and tilt of the instrument

housing. Air convection has been identified as a major source of

noise a long time ago. Another source of noise on seismographs

with helical and leaf springs has recently been identified and stud-

ied by Forbriger (2007): variations of Earth’s magnetic field. We

do not consider these effects here because we are interested in the

noise which cannot be avoided by clever construction, shielding and

installation of the sensors.

Tilt contributions to the noise in the free mode band associ-

ated with changes of barometric pressure were already studied by

Große-Brauckmann (1979) in the context of borehole installations

780 C© 2007 The Authors

Journal compilation C© 2007 RAS

Administrator
Highlight
Administrator
Highlight
Administrator
Highlight
Administrator
Highlight
Administrator
Highlight
Administrator
Highlight
Administrator
Highlight
Administrator
Highlight
Administrator
Highlight
Administrator
Highlight
Page 2: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

Horizontal seismic noise from barometric pressure 781

of Askania tiltmeters in Germany. Such contributions were ef-

fectively reduced in horizontal seismograms by Beauduin (1996),

Beauduin et al. (1996) and Neumann & Zurn (1999) in order to

improve signal-to-noise ratios (SNR). It is highly desirable to un-

derstand the physical phenomena producing disturbances in seismo-

grams because such understanding could lead to technical measures

or correction algorithms to reduce this noise. Of course, this is also

true for noise caused by atmospheric phenomena. Especially the

higher noise level in horizontal seismograms has led to a high im-

balance in the scientific use of vertical versus horizontal component

records at long periods. In this paper we describe simple models for

atmospheric disturbances and test their efficiency in noise variance

reduction. These physical models are intentionally kept simple with

only one free parameter for each of them for two reasons. The first

reason is the success of the similarly simple model for reduction of

vertical seismic noise below 1–2 mHz described by Zurn & Widmer

(1995). The second reason is that we would like to first learn about

the performance of simple models und gain insights into the phys-

ical mechanisms of tilt noise generation before more complicated

models with more parameters are applied.

A horizontal accelerometer is sensitive to several physical ef-

fects associated with a certain geophysical phenomenon. The iner-

tial effect (d’Alembert force) due to its proportionality to frequency

squared clearly dominates for high frequency signals and most of

body wave seismology. Varying tilts couple a varying component of

the gravitational acceleration g0 into the sensitive direction of the

seismograph and since g0 is very large compared to seismic accel-

erations away from the near-field of strong earthquakes, small tilt

angles produce significant disturbances. Mass redistributions due

to deformations produce varying horizontal gravitational forces on

sensor masses. In special cases the source of a phenomenon has a di-

rect gravitational effect on the sensor (like moon and sun in the case

of the Earth tides). The relative contributions of these four effects

to certain signals at different frequencies are listed in Table 1. Other

effects exist but are usually negligible. The four effects are com-

pletely independent of the construction details of the sensors. An

extensive and enlightening discussion of sensor sensitivities for hor-

izontal seismographs can be found in Rodgers (1968) who investi-

gates special construction features leading to systematic distortions

of seismic signals. Most if not all of those are avoided in mod-

ern broad-band force-balance sensors (e.g. Wielandt & Streckeisen

1982; Wielandt 2002).

Pendulum tiltmeters and pendulum horizontal seismographs are

in principle sensing the same effects, the distinction in the nomencla-

ture arises mainly from the different intentions of the deployers and

the frequency range where signals are expected. The idea of mea-

Table 1. Contributions (in per cent of the total signal) by relevant forces of

phenomena sensed by horizontal accelerometers. The amplitudes can only

be compared within one line. The increasing importance of the inertial effect

with rising frequency is conspicuous. For the tide the contributions to the

diminishing factor are listed. The data for the free modes were taken from

Dahlen & Tromp (1998, table 10.1). 1 S1 is the Slichter mode.

Signal Frequency Inertial Tilt Mass redistr. Gravitation

M 2-tide 22.4 μHz 0.6 −86.2 43.1 143.7

1 S1 0.0513 mHz −2.8 101.9 0.9 0.0

0 S2 0.3093 mHz −12.1 214.8 −102.7 0.0

0 S6 1.0382 mHz 81.0 21.4 −2.4 0.0

3 S8 2.8196 mHz 99.0 1.0 0.1 0.0

Rayleigh-wave 0.05 Hz 99.00 1.0 0.0 0.0

P wave 1.0 Hz 100.00 0.0 0.0 0.0

suring only tilts with a tiltmeter and the combination of horizontal

inertial accelerations and tilts with a horizontal accelerometer and

to subsequently use the two records to isolate the inertial accelera-

tion and to thus remove the tilt noise from seismograms (or similar

schemes) is as old as the knowledge about the tilt noise. However,

this idea is fallacious because tiltmeters also sense the inertial ac-

celerations. This applies not only to pendulum tiltmeters but also

to tiltmeters employing some form of liquid level (e.g. accelerate

a full cup of hot coffee horizontally without tilting it and a painful

experience will be the result). Examples for this are presented by

Ferreira et al. (2006) using records of the free oscillations excited

by the great 2004 N. Sumatra–Andaman Islands quake from a 40 m

water-tube tiltmeter in Luxembourg and from a 120 m differential-

pressure fluid-tiltmeter at Black Forest Observatory (BFO, see be-

low). Comparing the data with predictions demonstrates that these

tiltmeters sense predominantly the inertial accelerations associated

with the toroidal modes, which on a spherically symmetric earth do

not cause tilts. One may devise a scheme to remove the tilt noise in

a certain frequency band by a combination of some ‘tiltmeter’ and

some ‘inertial seismometer’, however, this will result in removing

the signals in this band as well.

In Section 2, we will describe the experimental determination of

the transfer-function between local barometric pressure variations

and seismograms. We will develop two simple physical models for

the atmospheric effects on seismograms in Section 3 and will com-

pare those models with long data sets in the following section. In

Section 5, the statistics for multiyear data sets are presented. Im-

provement of SNR is evaluated for a few examples in Section 6 and

work with finite-element models (FE) is briefly outlined in Section 7.

The resulting regression coefficients are discussed in Section 8 and

conclusions are drawn in the final section.

2 E X P E R I M E N TA L D E T E R M I N AT I O N

O F T R A N S F E R - F U N C T I O N

All our examples are observed at the GSN-station BFO (48.33◦N,

8.33◦E, 589 m amsl). The high quality of BFO is mentioned sev-

eral times in the literature (e.g. Beauduin et al. 1996; Freybourger

et al. 1997; Lambotte et al. 2006). This observatory is installed in

an abandoned silver–cobalt ore mine excavated into the surrounding

granite. The overburden above the instrument vaults is 150–170 m.

All seismographs at BFO are installed behind an efficient airlock

at a horizontal distance of ≈400 m from the adit. This airlock acts

as a low-pass filter for barometric pressure variations. Richter et al.(1995) give it’s time constant as ≈4 hr, this was increased to ≈36 hr

since then by improving the sealing. Thus pressure variations in the

long period seismic band are efficiently reduced in the galleries and

vaults behind the airlock and hence in the vicinity of the sensors.

The airlock also reduces temperature fluctuations to less than about

3 mK and prevents air circulation to a large extent. Small-scale con-

vection could still exist due to the little heat created by instruments

and electronics. Barometric pressure variations at BFO are sensed

by a mechanical microbarograph equipped with an inductive dis-

placement transducer. This unit has a sensitivity of 43.8 mV hPa−1

as calibrated by comparison with a precise mercury barometer and

a Paroscientific Digiquartz 740-16B. It is installed in the mine for

temperature stability but in front of the airlock to measure outside

pressure variations.

We attempted to estimate experimentally the transfer-functions

at BFO between local air pressure (outside of the airlock) and hori-

zontal accelerations using data from the Streckeisen STS-1/NS and

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Administrator
Highlight
Administrator
Highlight
Page 3: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

782 W. Zurn et al.

10 102

101

102

Pa >>> STS 1/NS Feb 1 28, 2005

Tra

nsfe

rF

un

ctio

n M

ag

nitu

de

[

nm

/ s

2 / h

Pa

]

Frequency [ Hz ]

TWM

LDM

103

102

0

0.5

1

Coherency

Figure 1. Experimental (noisy solid line) admittance between local air pres-

sure and NS-acceleration versus frequency for data from the STS-1 seis-

mometer at BFO from 2005 February 1 to 28, inclusively. Admittances for

the models developed in Section 3 are induced [LDM with 30 nm s−2 hPa−1

(broken line) and TWM with elastic μ = λ = 25 GPa and ch = 10 ms−1].

Insert shows coherency versus frequency.

EW-seismometers (all these data are disseminated through

IDA/IRIS). Two continuous time-series from each of the three in-

struments sampled with a rate of 0.2 Hz from February 1–28, and

March 1–27, 2005 were analysed as follows. These time-series were

chosen because during these 2 months nobody entered the mine

through the airlock (the earthquake activity during these two pe-

riods is described in Sections 4.2 and 4.3, respectively). The seis-

mograms were detided while earthquake signals and other distur-

bances were left in the records (see Figs 13 and 16). The mean

values were removed from all series. Segments overlapping by 7/8

with 8192 points (≈11.4 hr) were multiplied by a Hanning window

and Fourier transformed. Auto and cross power spectral densities

were computed and averaged over all these windows resulting in

estimates of the transfer-functions and coherencies as functions of

frequency. The admittances were corrected for the instrumental re-

sponses to represent the admittances between barometric pressure

and horizontal accelerations and were further smoothed over 9/10

of a decade. The results for the two components and the two time

windows are depicted in Figs 1–4. Inserts show the corresponding

10 102

101

102

Pa >>> STS 1/EW Feb 1 28, 2005

Tra

nsfe

rF

un

ctio

n M

ag

nitu

de

[

nm

/ s

2 /

hP

a ]

Frequency [ Hz ]

TWM

LDM

103

102

0

0.5

1

Coherency

Figure 2. Same as Fig. 1 for EW-component. Model admittances are now

15 nm s−2 hPa−1 for LDM and for TWM μ = λ = 60 GPa and ch =10 ms−1 were used.

10 102

101

102

Pa >>> STS 1/NS Mar 1 27, 2005

Tra

nsfe

rF

un

ctio

n M

ag

nitu

de

[

nm

/ s

2 /

hP

a ]

Frequency [ Hz ]

TWM

LDM

103

102

0

0.5

1

Coherency

Figure 3. Same as Fig. 1 for the NS-data from 2005 March 1 to 27, including

models.

10 102

101

102

Pa >>> STS 1/EW MAR 1 27, 2005

Tra

nsfe

rF

un

ctio

n M

ag

nitu

de

[

nm

/ s

2 /

hP

a ]

Frequency [ Hz ]

LDM

TWM

103

102

0

0.5

1

Coherency

Figure 4. Same as Fig. 3 for EW-component. Model parameters as in Fig. 2.

coherencies versus frequency. These have to be interpreted with cau-

tion (optimistically) because the arrivals of earthquake signals and

occasional disturbances of instrumental origin will always cause a

drop in the coherency with pressure. There appears to be a plateau

in the admittances up to about 1 and 0.6 mHz for the NS- and EW-

components, respectively. Above these frequencies the admittances

drop into a wide trough with the minima being about half the value

of the plateaus and then they rise steeply near about 8 mHz. How-

ever, this rise is probably not significant since the coherency is very

low. The plateau for the NS-component is near 30, for EW near 15

nm s−2 hPa−1.

In order to possibly improve the significance of the admittance

estimates we selected an interval from the first of the above time-

series which had only minor seismic activity and strong barometric

pressure variations, from 2005 February 10 to 13 and performed

an identical analysis. The results are shown in Figs 5 and 6. The

coherencies are much higher now. The plateaus are at similar values

as before but extend to higher frequencies, the troughs are shallower

(NS) or almost non-existent (EW), and the rises above 8 mHz are

much less pronounced.

Lambotte et al. (2006) also estimated admittances between lo-

cal barometric pressure and horizontal accelerations as observed

with the STS-1 seismometers at BFO (Figs 11 and 12 of their pa-

per) without specifying the time period used. They filtered the data

between periods of 7 d (≈1.6 μHz) and 3 hr (≈9.3 μHz) before

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 4: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

Horizontal seismic noise from barometric pressure 783

10 102

101

102

Pa >>> STS 1/NS Feb 10 13, 2005

[ n

m /

s2 / h

Pa ]

Frequency [ Hz ]

TWM

LDM

103

102

0

0.5

1

Coherency

Figure 5. Same (NS) as Fig. 1 for the much shorter window 2005 February

10–13. Strong pressure variations occurred during these days. Model param-

eters here are 35 nm s−2 hPa−1 (LDM) and 25 GPa and 10 ms−1 (TWM).

10 102

101

102

Pa >>> STS 1/EW Feb 10 13, 2005

Tra

nsfe

rF

unction M

agnitude

[ nm

/ s

2 / h

Pa ]

Frequency [ Hz ]

TWM

LDM

103

102

0

0.5

1

Coherency

Figure 6. Same as Fig. 5 for EW-component. Model parameters:

20 nm s−2 hPa−1 for LDM; 60 GPa and 10 ms−1 for TWM.

their analysis; thus their results are only slightly overlapping in fre-

quency with this study at the low end. Their admittances for the hor-

izontal components show an extended plateau with values of about

35 nm s−2 hPa−1 for NS and 20 nm s−2 hPa−1 for EW from about

20 μHz to 1 mHz. These estimates are very close to the values of

our plateaus at low frequencies.

3 T W O S I M P L E M O D E L S O F

AT M O S P H E R I C E F F E C T S O N

H O R I Z O N TA L S E N S O R S

In the following two simple physical models are described which

could be responsible for producing seismic signals as a result of

changes in barometric pressure. Both models were already intro-

duced by Neumann (1997), Neumann & Zurn (1999) and Zurn &

Neumann (2002). Zurn et al. (1999) utilized these concepts to re-

duce noise in the free mode records from BFO of the 1998 Balleny

Islands earthquake (M w 8.2). Preliminary tests on longer time-series

were described briefly by Exß & Zurn (2002).

The first model is based on observations and modelling of tilts and

strains caused by Earth tides: the local deformation model (LDM).

The second model was developed because wave-like variations in

atmospheric pressure have clear counterparts in horizontal (and ver-

tical) seismic records (e.g. Muller & Zurn 1983; Tsai et al. 2004;

Kroner et al. 2005; Zurn & Wielandt 2007): the travelling wave

model (TWM). Since the propagation velocity of atmospheric waves

is most often less than the speed of sound c ≈ 330 ms−1, that is, small

compared to seismic velocities in rocks which are of the order of

several km s−1, we treat the deformation in the crust quasi-statically

in both models.

If we are able to formulate the change in a force on the sensor mass

in terms of a change in local barometric pressure we can formally

produce ‘pressure seismograms’, which in turn can be compared to

recorded seismic traces. In the frequency domain we write

s(ω) = Ha(ω)

[δa

δp(ω)

]�p(ω), (1)

where H a is the transfer-function of the seismometer with respect

to ground acceleration and ω is the angular frequency. δa/δ p is the

proportionality factor or admittance of the physical effect producing

an acceleration of the sensor mass due to the pressure change, and

s and �p are the Fourier transforms of the seismogram and the

pressure variations, respectively.

3.1 Local deformation model (LDM)

Observations of tidal strains and tilts often show significant local

variations in amplitude and phase (e.g. Baker & Lennon 1973). This

very disturbing result became understood in the 1970s in terms of

local modifications of the tidal deformation field due to lateral het-

erogeneities (cavities, topography, geology; see Harrison 1985, for

a summary). King et al. (1976) suggested that these local modifica-

tions will occur for seismic displacement fields as well. Beauduin

et al. (1996) found different reactions of STS-1 horizontal seis-

mometers to barometric pressure variations on two different piers

within the same station. This is a clear manifestation of ‘cavity’

effects, the deformation field including tilts produced by the baro-

metric pressure load and thus the seismometer’s reactions are lo-

cally highly variable in space. Neumann & Zurn (1999) invoked

such effects in order to explain barometric signals on horizontal

seismograph and strainmeter records in addition to direct effects

from an acoustic-gravity wave (AGW) model as described in the

next subsection. Kroner et al. (2005) and Steffen et al. (2005) used

this (LDM) concept in FE models of the cavities, topography and

geology at stations BFO and Moxa, Germany, to explain the ef-

fects of observed transients in barometric pressure on horizontal

seismometers and strainmeters.

For the LDM we assume a horizontally homogeneous atmosphere

having only arbitrary density profile with height which is a function

of time. The corresponding surface pressure p then represents a

varying load on the surface and varying deformation of the crust

will result. If we assume also the rocks to behave in an elastic linear

fashion, then displacements, tilts, strains, etc. will be proportional

to the pressure variation, no matter how complicated the geology,

topography and cavity geometry may be at the site under considera-

tion. If anelasticity is negligible the variations are either perfectly in

phase or 180◦ out of phase with the pressure variations. Since here

we are concerned with accelerations, a, we write

�a(x, y, z, t) = C(x, y, z)�p(t), (2)

where C is a real constant depending on exact location and on com-

ponent but not on time. Similar equations can be written for dis-

placements, strains, tilts or rotations. In particular, if two orthogo-

nal horizontal components are affected, the direction of the apparent

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 5: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

784 W. Zurn et al.

acceleration is constant as well completely independent of the di-

rection of travel of the (very slowly) moving air masses. Indeed,

Beauduin (pers. comm. 1996) found a very stable direction of po-

larization for the pressure correlated horizontal signals discussed by

Beauduin et al. (1996).

3.2 Travelling wave model (TWM)

According to Hines (1972) AGW are ubiquitous in the atmosphere.

Sorrells (1971) calculated the effect of atmospheric pressure waves

on seismograms, but since he was concerned with higher frequencies

he considered the inertial effect only. Neumann (1997) and Neumann

& Zurn (1999) were the first to our knowledge to use the model of

AGW to study observed pressure effects in long period seismograms

and straingrams. They were able to model two observed signals

with the inertial, gravitational and tilt effects computed from the

simultaneous barometric pressure record. Mass redistribution was

neglected (for a discussion of this effect in terms of the vertical

component see Zurn & Wielandt 2007). Neumann & Zurn (1999)

found that in these examples significant contributions corresponding

to the LDM had to be added to the ones from TWM in order to get

close to the observations in magnitude and phase. In contrast, the

simultaneous strain records could be very well modelled using LDM

only, that is, strain signals were nearly in phase with the barometric

pressure variations.

Our TWM consists of an atmosphere overlying a homogeneous

elastic half-space. The surface of the half-space is the x,y-plane

with the z-axis pointing upward. The Lame parameters of the half-

space are λ and μ, but in all applications we let λ = μ. The scale

height of the atmosphere is H = c2/(γ g0), where c is the sound

velocity, g0 the surface gravity and γ = c p/cv , the ratio of specific

heats at constant pressure and volume, respectively. It can be shown

(e.g. Gossard & Hooke 1975; Neumann 1997; Nappo 2002; Nishida

et al. 2005) that the atmosphere supports AGW. If such a wave

travels along the surface and its amplitude decays exponentially

with height z it is called a Lamb wave and its density distribution

can be described by

ρ(x, y, z, t) = ρ exp

(− z

)exp[i(khx − ωt)], (3)

where ρ is the peak density at the bottom of this model atmosphere,

k h the horizontal wavenumber, i = √−1, and ω the angular fre-

quency. The resulting pressure at the surface has the same distribu-

tion over the surface and in time with

p(x, y, 0, t) = c2ρ. (4)

The horizontal phase velocity is ch = ω/k h, in the case of Lamb

waves close to the speed of sound c (Gossard & Hooke 1975). A

Lamb wave is well described by these equations. For a general

AGW only the loading effects but not gravitational attraction are

described well since it travels at an angle to the surface and the den-

sity distribution is more complicated. The AGW obey the dispersion

relation

k2z = ω2 − ω2

0

c2+

(ω2

B

ω2− 1

)k2

h, (5)

with ω0 = c/(2H ) being the acoustic cut-off frequency and ωB =√g/H − g2

0/c2 the Brunt-Vaisala frequency of the atmosphere

(e.g. Nappo 2002; Jones 2005). k z is the vertical wavenumber com-

ponent.

The horizontal phase velocities of AGWs range mostly between 5

and 330 ms−1 (higher horizontal phase velocities are possible, when

a high total phase velocity AGW travels under a steep angle with the

surface) and are thus much slower than seismic waves with speeds

of a few km s−1. Therefore, inertial terms in the elastic equations

for the solid half-space are not included, that is, the moving pressure

wave does not act as a source of seismic waves but deforms the half-

space quasi-statically. This is clearly an approximation in our model.

The sensors are located at x a , ya = 0 and za = 0. The right-hand

sides of the eqs (7)–(12) have to be multiplied by the factor

exp[i(khxa − ωt)]. (6)

With G denoting Newton’s gravitational constant the result for

the horizontal component of the Newtonian attraction of the sensor

mass by the air is (Neumann 1997)

�aNh (xa, 0, 0) = i2π

G

g0

p1

1 + ωc2

chg0

. (7)

The displacements are calculated following Sorrells (1971) for

the quasi-static case and the following terms are derived. The ap-

parent horizontal acceleration due to tilt is

�aTh (xa, 0, 0) = i

g0

λ + 2μ

λ + μp (8)

and the inertial effect from the horizontal acceleration is

�aIh (xa, 0, 0) = i

1

μ

λ + μωch p. (9)

The tilt is independent of frequency but phase shifted by 90◦

with respect to air pressure. This signal, therefore, corresponds to

the Hilbert transform of the air pressure. Mockli (1988) studied

horizontal noise at a temporary station at Zurich, Switzerland and

found a remarkable similarity of the air pressure induced signals to

the Hilbert transform of pressure. The corresponding expressions

for the vertical accelerations for the TWM can be found in Zurn

& Wielandt (2007). The three terms aNh , aT

h and aIh from eqs (7)

to (9) are in phase as can be simply understood using one of the

‘Gedankenexperiments’ in Zurn & Wielandt (2007), which we re-

peat here. Consider a vertical plane through the atmosphere above

a horizontal seismometer sensing accelerations perpendicular to it.

Now let the atmosphere to the right-hand side of the plane change

its density sinusoidally with time. The Newtonian attraction of the

sensor mass will be maximum to the right-hand side when the den-

sity has its maximum. The downward deflection of the surface due

to pressure loading will be maximum at the same instant, thus the

tilt will be to the right-hand side as well and the induced compo-

nent of the earth’s gravitational acceleration will also pull the sensor

mass to the right-hand side. The horizontal ground displacement at

the same instant will be maximum to the right-hand side, the corre-

sponding ground acceleration will be to the left-hand side and the

d’ Alembert force on the sensor mass to the right-hand side. Hence

all three contributions are in phase.

To facilitate the comparison with the LDM we also give the strains

�εxx (xa, 0, 0) = − 1

λ + μ

)p, (10)

�εzz(xa, 0, 0) = − 1

λ + μ

)p, (11)

and

�εzx (xa, 0, za) = 1

μkh za p, (12)

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 6: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

Horizontal seismic noise from barometric pressure 785

which vanishes for za = 0 as it should at the surface to satisfy

the boundary condition. The strains are in phase with the pressure

variations (and thus with the LDM accelerations) and 90◦ out of

phase with the horizontal accelerations. According to Farrell (1972)

for any pressure distribution p(x , y) over an elastic half-space the

areal strain is �ε areal = −1/2 p/(λ + μ), where p is the pressure.

This relation also holds in the case here (ε yy = 0).

The total admittance for a horizontal sensor for TWM is thus(�ah

�p

)= i

(2πG

g0

1

1 + ωc2

chg0

+ g0

λ + 2μ

λ + μ

)

+ i

(1

μ

λ + μω ch

)(13)

The terms are the gravitational attraction, the tilt and the inertial

effects on the sensor mass. We can relate these terms to each other

tilt

inertial= g0(λ + 2μ)

μωch

(14)

and

tilt

gravitation= λ + 2μ

2μ(λ + μ)

g20

2πG

(1 + ωc2

chg0

). (15)

Fig. 7 shows the magnitudes of the transfer-function as functions

of frequency for two different parameters μ = λ and three values

of ch. Fig. 8 depicts the ratios of the tilt effects to the gravitational

and inertial effects, respectively, as functions of frequency for two

extreme horizontal phase velocities. Since tilt dominates over most

of the frequency range we later use only the tilt contribution (TWT)

for the AGW pressure seismograms.

It is clear that all forces from the TWM are perpendicular to the

wave crests. Therefore, for a certain accelerometer the right-hand

side of eq. (13) has to be multiplied by the cosine of the angle

between its sensitive direction and the direction of propagation of

the plane AGW, that is, by numbers between −1 and +1. Unless

predominant directions of propagation of atmospheric waves exist

at a certain station, for extended time-series a stable experimental

admittance cannot be expected simply for this reason.

10 105

104

103

102

101

101

102

Transfer Function Pa >>>> a

h

ma

gn

itu

de

[ n

m /

s2 /

hP

a ]

Hz

20 GPa

100 GPa

TWM

Figure 7. Magnitudes of the transfer-functions between local barometric

pressure and horizontal acceleration for the TWM. Solid, dotted, and dashed

lines were calculated for ch = 10, 80 and 330 ms−1, respectively. Two values

of λ = μ were used as indicated.

10–6

10–5

10–4

10–3

10–2

10–1

100

101

102

Traveling Wave Model: Transfer function Pa >>> a

h

Magnitude [

nm

/ s

2 /

hP

a ]

Hz

Total

Grav.

Inertial

Tilt

TWM

300 m/s

10 m/s

Figure 8. Magnitudes of the transfer-functions between local air pressure

and horizontal accelerations for the TWM. Total effect and the three contri-

butions are shown for two values of ch: 10 and 300 m s−1. λ = μ = 50 GPa.

The tilt effect is identical for the two cases.

4 T E S T S O F T H E M O D E L S

With a method for reduction of horizontal noise in mind similar to

the one efficiently used for long period vertical noise (e.g. Zurn &

Widmer 1995) we performed experiments with seismograms from

the horizontal STS-1 seismometers at BFO. Following eq. (1)

we computed ‘pressure seismograms’ using the simultaneously

recorded local barometric pressure variations for the LDM and for

the tilt (TWT) part of the TWM. For the TWM tilt is the dominant

contribution (see Fig. 8) and also does not involve the horizontal

phase velocity ch which in reality is probably highly variable in

time and cannot be determined without an array. In addition, pre-

liminary tests with the data used in Section 4.1 showed that the

other two terms in eq. (13) do not contribute significantly to the

variance reductions. Of course, the direction of atmospheric wave

propagation is also highly variable. In the pressure seismograms

we did not specify the constants in eqs (2) and (8), the coefficients

found for the various analyses are discussed collectively in terms of

these constants in Section 8 below. Thus the horizontal accelerations

representing LDM and TWT were proportional to the pressure vari-

ations and their Hilbert transform, respectively. The consequence

is that the two model seismograms are orthogonal and the result-

ing model admittances, therefore, will not depend on whether both

models are fit simultaneously or individually.

Three different time-series were selected during which the airlock

was not opened. The raw seismograms for NS- and EW-components

sampled at 5 s intervals were first decimated after appropriate low-

pass filtering to a sampling interval of 20 s and then detided. Since

we intended to concentrate on the long period part of the seismic

spectrum we subsequently low-pass filtered the seismograms with

a cut-off frequency of 10 mHz. The mean was removed from the

barometric pressure series, which was then decimated to 20 s sam-

pling, low-passed with cut-off of 10 mHz using the identical filters

as for the seismograms and Fourier transformed. The spectrum was

multiplied by the complex transfer-function of the STS-1 with re-

spect to acceleration according to eq. (1). This results directly in the

spectrum of the pressure seismogram for the LDM and multiplica-

tion by i = √−1 gives the one for the TWT, except for the constant

real factors. Inverse Fourier transform of the spectra provides the

corresponding ‘seismograms’. Then one full day of data at either

end of the time-series was removed to avoid edge effects completely.

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 7: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

786 W. Zurn et al.

For different time windows within the different time-series least-

squares fits of the two ‘pressure seismograms’ (together with a con-

stant accounting for offsets) to the horizontal real seismograms were

performed. This results in constant coefficients for each component

and model: cNSLDM, cEW

LDM, cNSTWT and cEW

TWT which adopt positive values

when tilts are to the N and E with increasing pressure. We also ob-

tain residual seismograms r NS and r EW and variance (σ 2) reductions

RNS and REW were computed

R = σ 2(s) − σ 2(r )

σ 2(s)× 100 per cent, (16)

where s represents the low-passed seismograms and r the corre-

sponding residual. It is important to realize that earthquake signals

and rare instrumental gliches were not removed from the seismo-

grams. However, in the second and third case those were clipped at

the top and bottom before the least-squares fits in order to reduce

the disadvantageous effects of large seismic amplitudes on the re-

sults. It has to be kept in mind though that the variance reduction

can never reach 100 per cent because of seismic arrivals, instrumen-

tal noise, and ultimately the incessant free oscillations of the Earth

first observed by Nawa et al. (1998) and probably excited by the

atmosphere and/or oceans (e.g. Kurrle & Widmer-Schnidrig 2006).

The latter have not been identified in horizontal seismograms yet

but must be present there with acceleration amplitudes of the order

of a few pm s−2, since this is their amplitude in vertical records.

The magnetic field induced noise discussed by Forbriger (2007) is

shown by that author to be not detectable in the horizontal STS-1

seismometers at BFO. However, the horizontal components of the

STS-2 at BFO (Section 5) clearly show this effect, not surprisingly

because of the leaf spring suspensions in that triaxial instrument

(see discussion in Forbriger 2007).

This imperfection of the models will also lead to biased estimates

of the coefficients (e.g. Draper & Smith 1966, chapter 2.12). An-

other kind of bias probably affects our coefficients: we realize that

we fit very noisy functions of time—the pressure seismograms—to

very noisy functions of time—the seismograms (e.g. Olsen 1998,

section 2.3).

4.1 July 2–9, 2000

This time-series was chosen because it was very recent when this

work was started and it did not contain very large earthquakes. Ac-

cording to the Harvard Catalogue of moment tensors the 6 (30)

largest events had moment magnitudes between 5.5 (5.0) and 5.9.

Therefore, the seismograms were not clipped before the least-

squares analysis. The original length was 10 full days starting at

0:00 hr UTC on 2000 July 1. Later on, the first and last day were

eliminated, as mentioned above. Fig. 9 shows the two seismograms,

the two model seismograms, and the barometric pressure. The model

seismogram for the LDM is practically a bandpassed version of the

barometric pressure itself. Inspection of these records shows that

the large disturbances in barometric pressure have their (modified)

counterparts in the pressure seismograms, of course, but also im-

pressively in the real seismic records. In addition, the seismic records

show a little earthquake activity and especially the NS-component

suffers from several one-sided instrumental gliches near the

beginning.

These data were then subjected to the least-squares fit of the two

models to each component in the time domain. First the total time-

series of 192 hr was used. The resulting variance reduction is shown

in Fig. 10 and the coefficients are listed in Table 2, case A1. Then

windows of lengths between 2 and 144 hr were used and shifted

1 2 3 4 5 6 7 8 9

0

0.02

0.04

0.06

0.08

BFO JUL 2 9, 2000 real and ’pressure seismograms

ST

S1

ou

tpu

t

[ V

]

d elapsed in month UTC

NS

EW

LDM

TWM

Pa

1 2 3 4 5 6 7 8 910

0

10

20

30

40

50

60

70

80

90

100

Pre

ssu

re

[ h

Pa

]

Figure 9. Detided and low-pass filtered seismograms in Volts from the STS-

1 horizontal seismometers at BFO for the indicated period of time; ‘pressure

seismograms’, and barometric pressure P a (right-hand scale). The pressure

seismograms were multiplied by a factor so their amplitudes have the order

of magnitude of the real seismograms.

100

101

102

0

10

20

30

40

50

60

70

80

90

100

ma

xim

al va

ria

nce

re

du

ctio

n

[ %

]

window length [ h ]

EW

NS

Figure 10. Maximum variance reductions obtained when fitting the LDM

and TWT pressure seismograms to the STS-1 data from 2000 July 2 to 9 for

windows of different lengths shifted through the data.

through the whole time-series. For window lengths shorter than

24 hr the time shift was 1 hr; longer windows were shifted by

1 d. The maximum variance reductions obtained for each case are

depicted in Fig. 10. Fig. 11 shows the variation with time of the

variance reductions and the coefficients for a window length of 8 hr.

Variance reductions vary between 0 and 95 per cent, hence the

method of noise reduction is often highly efficient in this time-

series. Comparison of the seismogram and the variance reductions

as functions of time in Fig. 11 suggests the conclusion that the

higher the noise the higher the variance reduction obtained. The

averages of the coefficients for all windows and for the windows

with variance reductions ≥40 per cent (81 and 93 out of 182 for

NS and EW) are listed in Table 2, cases A2 and A3. These num-

bers will be discussed in Section 8. The NS-component responds

more strongly to the LDM than to TWT, while the opposite is true

for the EW-component. The larger coefficients in both cases (cNSLDM

and cEWTWM) remain always negative, while the smaller two (cEW

LDM and

cNSTWM) oscillate more or less around zero. All coefficients appear

to be correlated or anticorrelated to the variance reductions. There

is a clear tendency for the coefficients to adopt larger magnitudes

when the variance reduction is high. This effect (the possible bias

mentioned above) is with high probability caused by the

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 8: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

Horizontal seismic noise from barometric pressure 787

Table 2. Coefficients found for different analyses and data. Fitting LDM and TWT to the seismograms of the three periods 2000 July

2–9 (A), 2005 February 2–27 (B), and 2005 March 2–26 (C) resulted in the first three groups. The magnitudes of the average coefficients

in all three cases are always larger when the windows with low variance reductions are excluded. Cases D1 and D2 show coefficients

found by minimizing spectral noise between low degree free modes for large earthquakes (see text) and D3, D4 and D5 are the values

found for a cold front and other pressure disturbances. Cases E refer to the results of the long-term performance study in Section 5 with

the numbers indicating the different seismometers (STS-0, STS-1, and STS-2) now. All coefficients in one column should be identical,

except for LDM in cases E0 and E2. The results from finite element models of BFO are given as cases F (see text). Positive (negative)

values represent tilts to the N (S) and E (W) for increasing barometric pressure. Units for all coefficients are 10−8 m s−2 hPa−1.

Case cNSLDM cNS

TWT cEWLDM cEW

TWT Remarks

July 2–9, 2000

A1 −2.60 −0.39 +0.85 −1.59 Total time-series

A2 −2.04 +0.13 +0.48 −1.23 Average: all windows

A3 −2.77 −0.34 +0.87 −1.61 Average: windows with R ≥ 40 per cent

A4 −3.21 −0.48 1.02 −1.92 Special 4-d window (Nr. 122)

February 2–27, 2005

B1 −1.51 +0.77 −0.37 −0.39 Total time-series

B2 −1.82 +0.42 +0.70 −0.00 Average: all windows

B3 −2.38 +1.37 +1.52 +0.25 Average: windows with R ≥ 40 per cent

B4 −1.42 +1.02 +0.20 −0.61 Special 4-d window

March 2–26, 2005

C1 −1.38 +0.66 +0.42 −0.38 Total time-series

C2 −1.80 +0.51 +0.68 −0.34 Average: all windows

C3 −2.93 +0.53 +1.26 −0.74 Average: windows with R ≥ 40 per cent

C4 −1.15 +0.94 +0.20 −0.38 Special 4-d window

Special events

D1 −3.50 +0.50 +1.00 −1.35 Balleny Islands earthquake 1998

D2 −3.35 +0.60 +0.80 −1.00 N. Sumatra–Andaman Islands earthquake 2004

D3 −2.62 +0.25 +0.40 −1.62 Cold front May 18, 2006

D4 −2.70 −1.91 +1.50 +0.04 Pressure waves January 27, 2006

D5 −3.08 1.03 +0.67 −2.14 Pressure wave train October 13/14, 2002

Long-term statistics

E0 2.39 −0.08 +1.50 −0.33 STS-0, ≥3 years

E1 −1.89 0.00 +0.74 −0.70 STS-1, 12 yr

E2 −1.40 +0.45 +0.21 −0.65 STS-2, 3 yr

Finite Element Computations

F1 −0.297 0.121 STS-1 pier, raw model, no airlock

F2 −0.211 0.108 STS-1 pier, refined model, no airlock

F2B −1.320 0.262 STS-1 pier, refined model, airlock

F3 1.113 −0.629 STS-0 pier, refined model, no airlock

F3B 2.240 0.671 STS-0 pier, refined model, airlock

imperfection of the model, that is (apparent) pressure variations

which do not have counterparts in the seismograms, and the noise

in the time-series.

Fig. 12 shows the result of the modelling for a section of the

seismograms in Fig. 9. We used the coefficients found in window 122

of Fig. 11 (Case A4, Table 2) and formed the residual seismograms

r NS and r EW because the variance reductions in this window were

among the highest obtained and appeared stable for the adjacent

windows. The noise reduction here is quite dramatic but probably not

all the meteorologically caused disturbances were removed. Even

if the two models were appropriate it could be that one or both

coefficients change with time within a given time window, but most

likely the models are too simple. Experiments utilizing the inertial

and gravitational accelerations from the TWM did not significantly

improve the fit.

4.2 February 2–27, 2005

The second (much longer) time interval was already used in Sec-

tion 2 (see Figs 1 and 2) to determine experimentally the transfer-

functions. Fig. 13 shows seismograms, pressure seismograms, and

pressure for this period. The Harvard catalog lists 1 event with mo-

ment magnitude 7.1 (February 5), 5 between 6.5 and 6.9, 10 between

6.0 and 6.4 and 26 between 5.5 and 5.9 in this time period. Hence

seismic activity was much stronger (number and magnitudes) in

this time-series than in the previous one. Therefore, the seismo-

grams were clipped at ±5 mV in this case. Otherwise the analysis

was carried out as before. The efficiency of our procedure was not as

good as in the previous case. In part this is certainly due to the larger

seismic activity in this time window. When the total time-series was

used for the fit the variance reductions were only 18.2 (NS) and

4.7 per cent (EW). The resulting coefficients are given in Table 2,

case B1. In a second analysis we used windows 8 hr long for the fit

and shifted those by 1 hr. Fig. 14 depicts the variance reductions and

coefficients obtained for the series of windows. Variance reductions

exceeding 40 per cent were obtained for only 31 (NS) and 21 (EW)

of the 616 windows. The averages of the coefficients for all windows

(case B2) and for the windows with variance reductions larger than

40 per cent (case B3) are also given in Table 2. The coefficients will

be discussed in Section 8.

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 9: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

788 W. Zurn et al.

1 2 3 4 5 6 7 8 9

0

5x 10

3 BFO JUL 2 9, 2000

ou

tpu

t [ V

]

EW

0 20 40 60 80 100 120 140 160 180100

0

100

R [

% ]

NS

EW

0 20 40 60 80 100 120 140 160 180

30

20

10

0

10

Co

eff

NS

0 20 40 60 80 100 120 140 160 180

30

20

10

0

10

Co

eff

EW

8 h window number

Figure 11. Variance reductions and model coefficients (in nm s−2 hPa−1)

for the NS- and EW-seismograms and the two models for the pressure effects

for the time-series from 2000 July 2 to 9 inclusively versus window number.

Time windows were 8 hr long and shifted by 1 hr. The variance reductions

for the EW-component are plotted negatively for reasons of clarity. The co-

efficients for the LDM are shown by solid lines, for the TWT by dotted lines.

The top trace shows the EW-seismogram from Fig. 9 for visual comparison;

the horizontal axis is given in elapsed days in July.

6 6.1 6.2 6.3 6.4 6.5 6.6 6.7 6.8 6.9 7

0

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

BFO July 2000

Ou

tpu

t sig

na

ls [

V ]

d elapsed in month (UTC)

NS

rNS

EW

rEW

Figure 12. Section of the seismograms and their residuals with strong baro-

metric effects and with high variance reductions (see Fig. 11). Constant

coefficients were used for the calculation of residuals (case A4, Table 2).

Fig. 15 shows that part of the time-series in more detail which

was used in Figs 5 and 6. The fit was performed over these 3 d.

The variance reductions obtained were 29 per cent (NS) and

12 per cent (EW) and the coefficients are given in Table 2, case B4.

This example shows that the rather high efficiency of the method

experienced for July 2000 cannot be expected all the time.

4.3 March 2–26, 2005

This long time-series was already used in Section 2. Seismograms,

‘pressure seismograms’, and pressure are shown in Fig. 16. The Har-

vard catalog lists one event with moment magnitude 7.1 (March 2),

2 between 6.5 and 6.9, 8 between 6.0 and 6.4 and 18 between 5.5

and 5.9. Hence seismic activity was only slightly less than in the

previous case and much higher than in the first case. Therefore, the

seismograms were clipped at ±4 mV (NS) and ±5 mV (EW) and

the same analysis as in the preceding section was carried out. When

the total time-series was used for the fit the variance reductions

5 10 15 20 25

0

BFO FEB 2 27, 2005 real and pressure seismograms

ST

S1

ou

tpu

t

[ V

]

d elapsed in month UTC

NS

EW

LDM

TWM

Pa

5 10 15 20 2520

0

20

40

60

80

100

120

Pre

ssu

re

[ h

Pa

]

Figure 13. Detided and low-pass filtered seismograms in Volts from the

STS-1 horizontal seismometers at BFO for the indicated period of time;

‘pressure seismograms’, and barometric pressure P a (right-hand scale). The

pressure seismograms were multiplied by a factor so their amplitudes have

the order of magnitude of the real seismograms.

5 10 15 20 25

0

5x 10

3 BFO FEB 2 27, 2005

outp

ut [ V

]

EW

0 100 200 300 400 500 600

50

0

50R

[

% ]

NS

EW

0 100 200 300 400 500 600

20

0

20

Coeff

NS

0 100 200 300 400 500 600

20

0

20

Coeff

EW

8 h window number

Figure 14. Variance reductions and model coefficients (in nm s−2 hPa−1)

for the NS- and EW-seismograms and the two models for the time-series

from 2005 February 2 to 27. Windows 8 hr long and shifted by 1 hr were

used. The variance reductions for the EW-component are plotted negatively

for reasons of clarity. The coefficients for the LDM are solid, for the TWT

dotted lines. The top trace shows the EW-seismogram from Fig. 13 for visual

comparison (horizontal axis is in elapsed days (UTC) in February).

10 10.5 11 11.5 12 12.5 13

0

0.005

0.01

0.015

0.02

0.025

0.03

0.035

BFO Feb 2005

Ou

tpu

t sig

na

ls [

V ]

d elapsed in month (UTC)

NS

rNS

EW

rEW

Figure 15. Section of the seismograms and their residuals with rather large

pressure variations.

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 10: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

Horizontal seismic noise from barometric pressure 789

5 10 15 20 25

0

0.01

BFO MAR 2 26, 2005 real and pressure seismograms

ST

S1 o

utp

ut [ V

]

d elapsed in month UTC

NS

EW

LDM

TWM

Pa

5 10 15 20 25

0

20

40

60

80

100

120

Pre

ssure

[h

Pa]

Figure 16. Detided and low-pass filtered seismograms in Volts from the hor-

izontal STS-1 seismometers at BFO for the indicated period of time, ‘pres-

sure seismograms’, and barometric pressure variations (right-hand scale).

The models were multiplied by a factor so their amplitudes resemble those

of the real seismograms.

5 10 15 20 25

0

2

4x 10

3 BFO MAR 2 26, 2005

outp

ut [ V

] EW

0 100 200 300 400 500

50

0

50

R

[ %

] NS

EW

0 100 200 300 400 500

20

0

20

Coeff

NS

0 100 200 300 400 500

20

0

20

Coeff

EW

8 h window number

Figure 17. Variance reductions and model coefficients (in nm s−2 hPa−1)

for the NS- and EW-seismograms and the two models for the time-series

from 2005 March 2 to 26. Windows 8 hr long and shifted by 1 hr were

used. The variance reductions for the EW-component are plotted negatively

for reasons of clarity. The coefficients for the LDM are solid, for the TWT

dotted lines. The top trace shows the EW-seismogram from Fig. 16 for visual

comparison (horizontal axis is in elapsed days (UTC) in March).

obtained were 12 per cent (NS) and 4 per cent (EW) and the co-

efficients given in Table 2 as case C1 were found. The variance

reductions and coefficients found when using windows of length 8

hr and shifted by 1 hr are shown in Fig. 17 as functions of the window

series together with the clipped EW-seismogram. The efficiency of

the method was even worse than in February 2005. The averages

of the coefficients for all windows (case C2) and for windows with

variance reductions larger than 40 per cent (case C3) out of a total

of 592 are presented in Table 2.

A special 3-d window with stronger barometric pressure varia-

tions was also analysed. Variance reductions were 20 per cent (NS)

and 5.4 per cent (EW) and the coefficients are presented as case

C4 in Table 2. Fig. 18 depicts the seismograms and their residuals.

Again the improvement is marginal and by far not as good as in

Fig. 12.

9 9.5 10 10.5 11 11.5 12

0

0.005

0.01

0.015

0.02

0.025

0.03

0.035

BFO Mar 2005

Ou

tpu

t sig

na

ls [

V ]

d elapsed in month (UTC)

NS

rNS

EW

rEW

Figure 18. Section of the seismograms of Fig. 16 and their residuals. The

coefficients are tabulated in Table 2, case C4.

5 L O N G - T E R M P E R F O R M A N C E

S TAT I S T I C S

After the successful application of the models to the seismograms

from July 2000 we decided to do a long-term study in order to shed

more light onto the method and its performance. Three different

modern broad-band seismometer sets were or are operating at BFO

and all those were installed in the seismic vault about 500 m hor-

izontally into the hill behind the airlock and at a depth of 170 m

below the surface. The walls of the nearly square vault are oriented

NS and EW. The STS-1s considered so far in this paper were in-

stalled in February 1993 and are still in operation. They are sitting

on a low concrete pier (surface about 20 cm above the floor) directly

cemented to the granite at the foot of the southern wall of the vault.

Before February 1993 prototypes of the STS-1 (STS-0 in the fol-

lowing) were installed on a pier with the surface about 80 cm above

the floor and cemented to the foot of the northern wall, this wall

itself and the eastern wall. Finally, an STS-2 is operating since 1990

on a second low pier at the southern wall only a few cms west of

the one with the STS-1s. This STS-2 is a part of the German Re-

gional Seismic Network (GRSN). The horizontal distance between

the piers with the STS-1s and the STS-2 is varying by up to 40 nm

with the Earth tides (Zurn et al. 1991). Since this is more than the

displacement expected from the theoretical body tide strain by a

factor of about 8 this is another manifestation of (apparent) strain

enhancement due to cavity effects.

We decided to use a window 4 d long (a typical window for com-

puting free mode spectra after large quakes) and shifted this through

the data sets in steps of 1 d. Otherwise the analysis followed the

one described and used in the previous section. The seismograms

were subjected to a simultaneous least-squares fit to the LDM and

TWT ‘pressure seismograms’ computed from the simultaneously

recorded atmospheric pressure variations. We ignored all distur-

bances and earthquakes, but tides were removed. Table 3 shows the

number of windows for each case for which variance reductions

above a certain lower cut-off value were obtained. Figs 19–24 show

the distributions of variance reductions (top panel) and coefficients

in bins of 0.05 × 10−8 m s−2 hPa−1 for the two models (lower two

panels). In the histograms for the model coefficients those windows

were excluded for which the variance reduction was ≤5 per cent;

the corresponding numbers can be found in Table 3.

Figs 19 and 20 show the results for the STS-1 seismometers for

the years 1993–2004. Clearly, for the majority of the 3376 win-

dows the variance reduction was less than 10 per cent. However, the

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 11: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

790 W. Zurn et al.

Table 3. Number of 4-d windows with variance reductions larger than a

cut-off value for the three data sets identified by the name of the instruments

involved.

STS-0 STS-1 STS-2Cut-off

per cent NS EW NS EW NS EW

0 771 771 3376 3376 643 643

5 479 473 1983 1298 367 211

10 424 397 1524 818 250 113

15 381 334 1211 540 155 61

20 330 271 940 351 94 37

25 293 233 662 219 61 24

30 257 185 455 138 42 19

35 218 137 293 87 35 14

40 185 94 194 55 18 8

45 145 68 126 41 15 4

50 113 39 85 26 11 3

60 58 7 34 11 4 1

10 20 30 40 50 60 70 800

20

40

60

80

100

# W

ind

ow

s

Variance reduction [ % ]

5 0 5

x 108

0

20

40

60

80

100

# W

ind

ow

s

Coefficient cLDM

NS5 0 5

x 108

0

20

40

60

80

100

Coefficient cTWT

NS

Figure 19. Histograms of variance reductions (top) and model coefficients

(bottom) for the data from the STS-1 seismometers at BFO for the time

period from 1993 to 2004. The coefficients were only adopted for windows

with variance reductions larger than 5 per cent (see numbers in Table 3).

Units for coefficients are in m s−2 hPa−1.

0 10 20 30 40 50 60 70 800

20

40

60

80

100

# W

ind

ow

s

Variance reduction [ % ]

5 0 5

x 108

0

20

40

60

80

100

# W

ind

ow

s

Coefficient cLDM

EW5 0 5

x 108

0

20

40

60

80

100

Coefficient cTWT

EW

Figure 20. Same as Fig. 19 for the EW-component of the STS-1.

0 10 20 30 40 50 60 70 800

5

10

15

20

25

30

# W

ind

ow

s

Variance reduction [ % ]

5 0 5

x 108

0

5

10

15

20

25

# W

ind

ow

s

Coefficient cLDM

NS5 0 5

x 108

0

5

10

15

20

25

Coefficient cTWT

NS

Figure 21. Same as Fig. 19 for the NS-component of the STS-0 for the time

period from 1990 to February 1993.

10 20 30 40 50 60 70 800

5

10

15

20

25

30

# W

ind

ow

s

Variance reduction [ % ]

5 0 5

x 108

0

5

10

15

20

25

# W

ind

ow

s

Coefficient cLDM

EW5 0 5

x 108

0

5

10

15

20

25

Coefficient cTWT

EW

Figure 22. Same as Fig. 21 for the EW-component of the STS-0.

0 10 20 30 40 50 60 70 800

10

20

30

40

# W

indow

s

Variance reduction [ % ]

5 0 5

x 108

0

5

10

15

20

# W

indow

s

Coefficient cLDM

NS5 0 5

x 108

0

5

10

15

20

Coefficient cTWT

NS

Figure 23. Same as Fig. 19 for the NS-component of the STS-2 from 2001

to 2003.

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 12: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

Horizontal seismic noise from barometric pressure 791

0 10 20 30 40 50 60 70 800

10

20

30

40

# W

indow

s

Variance reduction [ % ]

5 0 5

x 108

0

5

10

15

20

# W

indow

s

Coefficient cLDM

EW5 0 5

x 108

0

5

10

15

20

Coefficient cTWT

EW

Figure 24. Same as Fig. 23 for the EW-component of the STS-2.

maximum R obtained reached 80 per cent (NS) and 70 per cent

(EW). The coefficients cNSLDM, cEW

LDM and cEWTWT are very nicely dis-

tributed around central values of about −3 × 10−8, +1 × 10−8 and

−1.3 × 10−8 m s−2 hPa−1, respectively, but only the latter two distri-

butions have tails reaching values of the opposite sign as seen from

the maximum. cNSTWT looks like a bimodal distribution straddling zero

with the negative values being less frequent. Such behaviour is basi-

cally not unexpected for TWT since the projection of the horizontal

phase velocity vector for travelling waves onto the sensitive direction

may be positive or negative. Since the distribution for EW is one-

sided, one could speculate that the EW-component of the horizontal

phase velocity ch for TWT is most of the time positive, while for

NS it fluctuates around zero. Table 2 (case E1) shows the means of

the coefficients for all windows with variance reduction larger than

5 per cent.

Figs 21 and 22 show the results for the STS-0 seismometers for the

years 1990 to the beginning of 1993. Since these were on a different

pier, the LDM could easily produce other values and polarities.

The TWT is a model common for all installations within a station;

if the coefficients really are representing the assumed underlying

physics they should have similar distributions for all installations

and piers within one station. As seen in Table 3, the relative number

of windows with high variance reduction is higher than for the STS-

1s. This could be an effect of the installation but also of the different

time period. All four coefficients show broader distributions with the

LDM coefficients being clearly one-sided around larger mean values

and the TWT-coefficients for both components clearly straddling

zero. Table 2 (case E0) shows the means of the coefficients for

windows with variance reductions larger than 5 per cent.

Finally, Figs 23 and 24 depict the results for the STS-2 seis-

mometer of the GRSN for the years 2001–2003. cNSLDM and cEW

TWT

are now one-sided and both positive with rather broad distributions

around −1 × 10−8 m s−2 hPa−1. The coefficients cNSTWT and cEW

LDM

both straddle zero now. The time period is simultaneous to part

of the time-series for the STS-1s. In this case the relative num-

ber of windows with high variance reduction is smaller than in the

other two cases for both components. Table 2 (case E2) lists the

mean values for the windows with variance reductions larger than

5 per cent.

As stated before, the LDM model coefficients can vary largely

within very short distances, while the TWT model coefficients

should be the same for each component independent of the place of

installation within one station. Comparing the histograms for the 12

coefficients shows that these properties are approximately obtained

in this study. The histograms for the TWT model look somewhat

similar to each other, with the ones for cNSTWT centred near zero and

the ones for cEWTWT centred at negative values. The histograms for

cNSLDM and cEW

LDM show much more variation in the three cases.

6 I M P ROV E M E N T O F

S I G N A L - T O - N O I S E R AT I O I N

S E I S M O G R A M S

6.1 Free oscillation spectra

The two models LDM and TWT were used in attempts to reduce the

pressure generated noise in free oscillation spectra from horizontal

records in the frequency band containing the lowest degree modes.

Towards this end for a given horizontal record a grid search in the

space of LDM and TWT-coefficients was carried out (a least-squares

fit would also be possible) and the minimum of the sum of Fourier

amplitudes between the known mode frequencies was sought. Then

the coefficients for the minimum were used to form the ‘corrected’

time-series (residuals).

The first two examples are the NS- and EW-seismograms from

the STS-1 sensors at BFO after the earthquake at the Balleny Is-

lands on 1998 March 25 (M w 8.2). Time-series of lengths 36–

44 hr were found to provide reasonable raw data for application of

the procedure outlined above since these showed some of the lower

degree modes above the noise. The resulting coefficients are listed

in Table 2 (case D1). Figs 25 and 26 show the spectra computed

from raw and corrected data for the NS- and EW-seismograms. In

both cases the noise between the modal peaks is clearly reduced in

the corrected records. For the EW-record all modes show improved

SNR after the correction. For the NS-spectrum the SNR for 0T 4

(0.766 mHz), 0T 5 (0.929 mHz) and others is enhanced by the cor-

rection, however, in the case of 0T 3 (0.587 mHz) it appears that the

correction does not help at all. This does not depend on the length

of the time-series. One possible speculative explanation could be

that in the raw data the contribution to the Fourier amplitude at the

0.2 0.4 0.6 0.8 1 1.20

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2x 10 BFO STS 1/NS Balleny Islands 1998

FF

T, 40 h

hanned, lin

ear

mHz

NS corrected

NS raw

0S

2 0S

3 0S

4 0S

5 0S

6 0S

7

0T

2 0T

3 0T

4 0T

5 0T

6 0T

7

Figure 25. Spectra of time-series from the STS-1/NS seismometer at BFO

starting at 3:00:00.0 hr UTC on 1998 March 25 and 40 hr long computed

after multiplication with a Hanning window. Spectral peaks correspond to

free oscillations excited by the Balleny Islands quake (bottom: raw data, top:

data corrected using the models). Vertical lines indicate the frequencies of

fundamental spheroidal modes 0 S2 to 0 S7 (dotted) and fundamental toroidal

modes 0T 2 to 0T 7 (dashed). The reduction of the noise is obvious, but 0T 3

appears to be lost due to the correction.

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 13: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

792 W. Zurn et al.

0.2 0.4 0.6 0.8 1 1.20

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2x 10 BFO STS 1/EW Balleny Islands 1998

FF

T, 40 h

hanned, lin

ear

mHz

EW corrected

EW raw

0S

2 0S

3 0S

4 0S

5 0S

6 0S

7

0T

2 0T

3 0T

4 0T

5 0T

6 0T

7

Figure 26. Same as Fig. 24 for the EW-component of the STS-1 at BFO.

Noise is clearly reduced between the modes and some of those show im-

proved SNR after the correction.

peak frequency from the barometric pressure noise and from the

mode itself are nearly in phase and taking the barometer effect away

reduces the total amplitude and pushes the peak back into the noise.

This is always a possibility with such corrections, but then the peak

detection in the raw spectrum is somewhat dubious. Incidentally, a

related but contrasting example occurs in vertical component data

from the same quake and station as shown in Fig. 1 of Zurn et al.(2000). There the barometric pressure effect apparently pulls the

frequency of the Coriolis-coupled mode 0T 3 away from its correct

position in the spectrum and after the correction its location agrees

with the long-known frequency.

The third example is the NS-component record of the STS-1 at

BFO from the great N. Sumatra–Andaman Islands event of 2004

December 26. The coefficients were found as described above and

are listed as case D2 in Table 2. Fig. 27 depicts spectra of the raw

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.10

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6x 10 N Sumatra 26.12.2004 BFO STS 1/NS

line

ar

am

plit

ud

e,

96

hF

FT,

ha

nn

ed

mHz

raw data

corrected

oS

2 oS

3 oS

4 oS

5 oS

6

oT

2 oT

3 oT

4 oT

5 oT

6

2S

1

1S

2

Figure 27. Spectra computed from the 96 hr—record of the STS-1/NS seis-

mometer at BFO starting at 0:00:00.0 hr UTC on 2004 December 26 and

multiplied by a Hanning window. It shows peaks corresponding to the free

oscillations excited by the N. Sumatra–Andaman Islands quake. Raw data

are shown at the bottom, top panel shows the spectrum after the record was

corrected with a combination of the LDM- and TWT-models found by mini-

mizing the noise power between the lowest degree modes(case D2, Table 2).

Vertical lines, as labelled, correspond to the frequencies of fundamental

spheroidal modes (0 Sl , dashed) and fundamental toroidal modes (0T l , dot-

ted). 0 S2 appears above the noise in the corrected data. Immediately to the

right-hand side of 0T 2 two singlets of the widely split overtone 2 S1, well

excited by this event, stand out more clearly.

2 3 4 5 6 7 8 9 10 11 12

0

0.005

0.01

0.015

0.02

0.025

0.03

BFO May 18, 2006

ST

S1 o

utp

ut

[ V

]

h UTC

NS

rNS

EW

rEW

LDM

TWT

Pa

2 3 4 5 6 7 8 9 10 11 1225

20

15

10

5

0

Pre

ssure

[ hP

a ]

Figure 28. Cold front passage of 2006 May 18 at BFO. From top to bottom

are shown: barometric pressure (right-hand scale), ‘pressure seismograms’

for TWT and LDM, detided original and residual STS-1/EW-seismograms,

detided original and residual STS-1/NS-seismograms (all in Volts, left-hand

scale).

and corrected records. The noise after the correction is reduced

considerably and the SNRs of the modes clearly improved. The

football mode 0 S2 (0.309 mHz) shows up above the noise level in

the corrected data (the record length is too short to see the splitting

of this mode). The EW-record was also subjected to this procedure

but only marginal improvement could be obtained.

6.2 Other examples

On 2006 May 18 a cold front was passing over BFO with an over-

all pressure increase of about 3.5 hPa. Clear disturbances in the

broad-band seismograms were observed at this time. Fig. 28 shows

the barometric pressure variations, the corresponding ‘pressure seis-

mograms’ for the LDM- and TWT-models, the raw and residual hor-

izontal seismograms from the STS-1 seismometers. The two ‘pres-

sure seismograms’ were fit simultaneously to the data for the 10 hr

time-series. The noise reductions obtained are impressive, not only

during the cold front passage. Variance reductions of 69 per cent

(NS) and 71 per cent (EW) were obtained. The corresponding co-

efficients are listed as case D3 in Table 2.

On 2006 January 27 several disturbances in barometric pressure

including a quasi-periodic wave train after 10:00 occurred at BFO

with spectral energy within the passband of the STS-1 seismome-

ters. Fig. 29 shows the barometric pressure, the ‘pressure seismo-

grams’ computed for the LDM and TWT models and the detided

seismograms and their residuals. The variance reductions obtained

were 82.5 per cent (NS) and 57.2 per cent (EW) and the coefficients

found are listed as case D4 in Table 2. The reduction of noise is

conspicuous, especially for the NS-component.

In the night from 2002 October 13 to 14 pressure disturbances

occurred at BFO. An especially nice quasi-periodic wave train can

be observed in the pressure record in Fig. 30 between 23 and 24 hr

UTC with the counterparts in the ‘pressure seismograms’ and the

horizontal records from the STS-1s. Seismic traces were detided and

those and the pressure record were low-pass filtered with a cut-off

frequency of 10 mHz. Then the model-seismograms for LDM and

TWT were computed from the pressure record and fit simultaneously

to the 10 hr data shown in Fig. 30. The resulting coefficients are

presented as case D5 in Table 2 and the variance reductions were

82.9 per cent (NS) and 78.3 per cent (EW). The improvement of the

SNR is again quite dramatic, also around the earthquake signal near

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 14: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

Horizontal seismic noise from barometric pressure 793

5 6 7 8 9 10 11 12 13

0

2

4

6

8

10

12

14

16

18

x 103 BFO Jan 27, 2006

ST

S1

ou

tpu

t

[ V

]

h UTC

NS

rNS

EW

rEW

LDM

TWT

Pa

5 6 7 8 9 10 11 12 1310

8

6

4

2

0

Pre

ssu

re

[ h

Pa

]

Figure 29. On 2006 January 27 several pressure disturbances occurred at

BFO. A wave train can be observed between 10:00 and 10:30 UTC shortly

after a cold front like rise in pressure and two other larger transients can

be identified at 6:00 and 11:50 UTC. All these have their signatures in the

seismograms. From top to bottom are shown: barometric pressure (right-

hand scale), ‘pressure seismograms’ for TWT and LDM, detided original

and residual STS-1/EW-seismograms, detided original and residual STS-

1/NS-seismograms (all in Volts, left-hand scale).

18 19 20 21 22 23 24 25 26 27 28

0

2

4

6

8

10

12

14

16

18

x 103 BFO OCT 13 14, 2002

ST

S1

ou

tpu

t

[ V

]

h UTC from 0:00:00.0 on Oct 13

NS

rNS

EW

rEW

LDM

TWT

Pa

18 19 20 21 22 23 24 25 26 27 2812

10

8

6

4

2

0

Pre

ssu

re

[ h

Pa

]

Figure 30. From 2002 October 13 to 14 several wavelike pressure distur-

bances occurred at BFO. A nice quasi-periodic wave train can be seen in the

pressure record between 23 and 24 hr with corresponding counterparts in

the seismograms. From top to bottom are shown: barometric pressure (right-

hand scale), ‘pressure seismograms’ for TWT and LDM, detided original and

residual STS-1/EW-seismograms, detided original and residual STS-1/NS-

seismograms (all in Volts, left-hand scale). The two seismic signatures at

22:00 and 22:30 appear with much improved SNR in r EW but are almost

absent in r NS, so these waves were polarized in the EW-direction.

22 hr UTC caused by the Samoa Islands quake with M w = 6.0 and

origin time 20:55 hr UTC. When we performed the same analysis for

the extended records (October 12, 5:00 hr to October 14, 19:00) we

obtained variance reductions of ‘only’ 54.9 per cent (NS) and 35.4

per cent (EW). When this longer record was low-pass filtered with

a cut-off of 2 mHz, the corresponding numbers were 53.3 per cent

(NS) and 46.0 per cent (EW). The conclusion can be drawn that

shortening the time-series or narrowing the frequency range often

may result in improved performance of this noise reduction method.

However, this clearly will depend on the coherency of the pressure

seismograms with the real ones as a function of time and of the

spectral power and range of the pressure disturbances.

7 F I N I T E E L E M E N T M O D E L S

Steffen et al. (2005) subjected a FE model of the geology, topogra-

phy and cavity geometry of the station BFO and its neighbourhood

to different schemes of loading by the atmosphere. In particular, uni-

form pressure over the whole model and inside the galleries as well

as wind-pressure on the flank of the hill (striking NS) were applied.

The resulting coefficients for uniform pressure can be compared

to the cNSLDM and cEW

LDM as estimated from the data. They modified

this model in different ways in order to understand the sensitiv-

ity of the resulting pressure coefficients to different features of the

model. Their model coefficients for tilts produced by uniform pres-

sure for a position in the model corresponding to the location of the

STS-1 seismometers as closely as possible are listed as case F1 in

Table 2.

The FE coefficients (F1) are smaller than the observed ones by an

order of magnitude for the NS-component and also largely underes-

timated for the EW-component. Therefore, subsequently the model

was much refined by introducing details of the piers in the seismic

vault and it was found that these details play a significant role. The

new coefficients for the pier with the STS-1s are presented as F2 in

Table 2 (Steffen et al. 2005, point P8). The NS-coefficient is about

75 per cent of that of the coarser model while the EW-coefficient

is slightly smaller (90 per cent). These results are again smaller

than the observed coefficients by about an order of magnitude

and are, therefore, unsatisfactory. We conjectured that this large

discrepancy is due to the fact that in the FE-model the pressure

was also acting inside the galleries while in reality the pressure

variations inside the inner part of the mine are reduced by at

least a factor of 20 by the airlock in the frequency range under

investigation.

Therefore, new FE-calculations were performed with no air pres-

sure loading from inside the galleries (i.e. a perfect airlock). The

resulting coefficients for the pier with the STS-1s are presented as

case F2B in Table 2. For both components the discrepancy with the

observed admittances becomes much smaller. The remaining differ-

ences are not understood at present but are thought to be related to

structural details in the immediate vicinity of the pier not included

in the FE-model.

The pier on which the STS-0 was installed was also modelled and

resulted in coefficients given in Table 2 as cases F3 and F3B with

and without pressure loading from inside, respectively. These can

be compared to case E0 and again the agreement is very poor for

the case with pressure loading from inside. For both components the

magnitudes of the FE-estimates were too small by about 50 per cent

and for EW even the sign was different. For the case without internal

loading the agreement between data and model is significantly im-

proved. For the NS-component the difference amounts to 6 per cent

which is very good considering the experimental variabilities. For

the EW-component the signs now agree but the model coefficient is

only 40 per cent of the observed one.

Attempts to better understand these remaining discrepancies are

under way. An interesting observation here is that according to these

models the airlock at BFO increases the sensitivity of the piers

to pressure loading. This can be understood as the effect of the

missing pressure variations inside the galleries which to some extent

would reduce the overall deformation of the rocks. However, the

benefits of the airlock remain undisputed, since it acts to reduce the

pressure variations in the immediate environment of the sensors by

at least a factor of 20 in the long-period seismic band, stabilizes the

temperature inside the mine to a few mK (Richter et al. 1995), and

reduces air flow.

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 15: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

794 W. Zurn et al.

8 D I S C U S S I O N O F T H E M O D E L

C O E F F I C I E N T S

The coefficients cNSLDM, cEW

LDM, cEWTWT and cNS

TWT listed in Table 2 basi-

cally all have the same order of magnitude but scatter appreciably

from experiment to experiment in the detailed values. With the ex-

ception of cNSLDM they even change polarity. However, the overall pat-

tern of the coefficients (Table 2 and Figs 19–24) is approximately

consistent with our models. For the TWM these polarity changes

and variabilities can be expected because in a certain time-series

the corresponding averages of the parameters (ch, direction of prop-

agation) can change. The real atmosphere is of course much more

complicated than this simple model. However, for the LDM the co-

efficients should be very stable, since the resulting tilts should not

depend on the atmospheric phenomena. Because the seismic data

contain many truly seismic ‘disturbances’ which have nothing to

do with barometric pressure our estimated coefficients necessarily

must be biased in each case by different amounts. Therefore, good

stability in the experimental results cannot be expected even in the

case of the LDM.

The constant C(x , y, z) in eq. (2) can immediately be identified

with the coefficients cNSLDM and cEW

LDM for the two components. As-

suming the validity of the formula for pressure-induced strain from

Farrell (1972) or eq. (10), λ=μ, and strain–tilt coupling coefficients

of the order of 1 nrad/10−9 as estimated for tides (Emter & Zurn

1985) we can derive an effective shear modulus for the crust and

our phenomenon near BFO by μ = 1/4 g0/cLDM. For the extreme

value of 3.5 × 10−8 m s−2 hPa−1 for case D1 and the NS-component

we obtain as a minimum μ = 6.9 GPa. For the EW-component the

coefficients are smaller by factors of 2–5 and μ becomes corre-

spondingly higher. In summary, these values are at least providing

the right order of magnitude while being a little small. The dif-

ferences between the components must be generated by the local

heterogeneities, that is, different strain–tilt coupling coefficients.

For the TWT-coefficients we have the problem, that the direction

of propagation is unknown and with high probability not constant.

With all uncertainties in mind we can again determine the order of

magnitude of the effective shear modulus using eq. (8) and λ = μ.

We get for the maximum in Table 2 a minimum in μ of 49 GPa,

much more reasonable than the value for LDM.

9 D I S C U S S I O N A N D C O N C L U S I O N S

The comparisons of our two simple physical models to the data

from the broad-band seismometers at BFO obviously resulted in

vastly different results as far as variance reductions are concerned.

This variability is not at all expected for the LDM-model, while for

TWM due to temporal variations in the phase velocity and direction

of travel it is an inherent property of the model. Our fits suffer to a

presently unknown extent from the fact that the model functions are

imperfect and noisy. This could possibly cause some of the variabil-

ity. Another source of variability arises from the fact that the seis-

mic data obviously contain seismic signals and unfortunately also

instrumental disturbances unrelated to barometric pressure. While

this will definitely always bias low the obtained variance reductions

it will also cause variations in the pressure admittances as soon as

these additional signals are not perfectly orthogonal to the baromet-

ric pressure. However, the order of magnitude of all the coefficients

and the general behaviour of each individual coefficient are at least

approximately corroborating the physics of the models. Even in the

successful cases there appear to be barometrically caused effects

left in the residuals, therefore, improvements are probably possible.

Clearly both of our models are not able to accommodate nei-

ther frequency dependence (outside of bandpass filtering the time-

series) nor time dependence (outside of selection of specific time

windows). Since the optimal coefficients for reduction of noise

show this high variability one extension of the method could be

to determine the coefficients in a data-adaptive way but with some

smoothing criterion included and provisions made how to handle

seismic waves larger than the noise level. Special frequency depen-

dencies could be prescribed in order to get even more efficient mod-

els, however investigating such models will be a subject of further

research.

Beauduin et al. (1996) were quite successful using a spectral do-

main approach for their noise reduction technique. They determined

a complex transfer function by finding Fourier coefficients at many

frequencies which maximized the coherency between seismic noise

and the local barometric pressure, similar to the transfer functions

determined in Section 2 of our paper. Their model for correction

has many more coefficients than the ones used here. They give their

transfer functions in units of m s−1 hPa−1 instead of m s−2 hPa−1

so in order to compare their absolute values with our coefficients

we have to multiply theirs with ω. Their maximum coefficients are

factors of between 2 and 10 larger than the largest coefficient in

Table 2. This means that they observed larger disturbances than we

did for some physical reason, that is, their horizontal components

are much more sensitive to barometric pressure than ours at BFO.

In passing we note that their maximum values for the vertical com-

ponent are a little smaller (e.g. at 0.5 mHz) than the values obtained

for the well-sealed superconducting gravimeters (between 2.7 and

4.0 nm s−2 hPa−1) and the LaCoste-Romberg gravimeter at BFO

(3.5 nm s−2 hPa−1, Zurn & Widmer 1995). This could be due to ei-

ther a larger effect due to atmospheric loading at their station SSB

with respect to BFO or some instrumental effect opposing the effect

due to gravitational attraction.

Considering Beauduin’s (1996) experience with different instal-

lations at one site and realizing that our STS-1/NS-seismometer has

a rather large sensitivity to barometric pressure one is led to the idea

that within one observatory one could attempt to find the position

with the lowest coefficients experimentally by doing a study like

the one here for sensors at different places. However, in both cases

the corrections worked very well when the effects were large. So it

is quite uncertain, if an improvement is possible and, therefore, the

STS-1/NS seismometer at BFO will not be moved since in general

the data quality is among the very best in the world.

Another, often suggested improvement could possibly be obtained

by the deployment of a microbarograph array or network around the

station in order to get the gradient of the pressure as a function of

time and the phase velocity and travel direction of the atmospheric

waves or masses (Egger et al. 1993; Tsai et al. 2004; Nishida et al.2005; Riccardi et al. 2007). However, the mean distance between

barometers in such a network probably should not exceed a few

kilometres in order to be helpful in the frequency range considered

here.

Despite all these shortcomings we think that our extremely sim-

ple models are very successful considering the positive results in

the work described. The method can be applied at any station which

records local barometric pressure in addition to broad-band seis-

mographs. It is especially promising at times and places when and

where barometric disturbances are especially strong. The order of

magnitude of our coefficients is not at variance with the physical

parameters of the crustal properties of the Earth.

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 16: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

Horizontal seismic noise from barometric pressure 795

The results of the finite element calculations for BFO are in gen-

eral underestimating the coefficients. An interesting observation was

made with these models: the airlock at BFO possibly increases the

sensitivity to outside pressure loading for the horizontal compo-

nents. This can clearly be seen when comparing the coefficients for

cases F2 and F2B, and F3 and F3B in Table 2. However, even if this

should be the case the benefits of the airlock overwhelm this effect.

The larger effects in the NS-components are clearly reproduced in

the FE-model. The remaining discrepancies are not understood at

present. One possible conclusion could be that all details in geom-

etry/geology in the vicinity of the seismic vault which are not mod-

elled act to increase the displacements due to barometric pressure

loading.

A C K N O W L E D G M E N T S

Discussions with Erhard Wielandt, Thomas Forbriger, Rudolf

Widmer-Schnidrig, Heinz Otto, Peter Davis and Klaus Klinge were

very helpful. We appreciate comments by Cathy Constable and

Andreas Junge. Numerous suggestions by Gabi Laske (Ed.) and

the scrutinous review by Luis Rivera as well as the comments

of two anonymous reviewers helped to improve the manuscript.

Dirk Ansorge carefully read the manuscript. Financial support by

the “Deutsche Forschungsgemeinschaft” under grants number KR

1906/3-1,2 and WE 2628/1-1,2 is gratefully acknowledged.

R E F E R E N C E S

Agnew, D.C. & Berger, J., 1978. Vertical seismic noise at very low frequen-

cies, J. geophys. Res., 83, 5420–5424.

Baker, T.F. & Lennon, G.W., 1973. Tidal tilt anomalies, Nature, 243, 75–76.

Beauduin, R., 1996. Etude du bruit de fond sismique a l’aide des donnees

GEOSCOPE et des donnees de l’experience OFM/SISMOBS, These dedoctorat, Universite de Paris VII, 130 pp.

Beauduin, R., Lognonne, P., Montagner, J.P., Cacho, S., Karczewski, J.F.

& Morand, M., 1996. The effects of atmospheric pressure changes on

seismic signals or how to improve the quality of a station, Bull. seism.Soc. Am., 86, 1760–1769.

Berger, J., Davis, P. & Ekstrom, G., 2004. Ambient earth noise: a survey

of the global seismographic network, J. geophys. Res., 109, B11307,

doi:10.1029/2004JB003408.

Dahlen, F.A. & Tromp, J., 1998, Theoretical Global Seismology, Princeton

University Press, Princeton, New Jersey, 1025 pp.

Draper, N.R. & Smith, H., 1966, Applied Regression Analysis, J. Wiley and

Sons, New York, 307 pp.

Egger, J., Wamser, C. & Kottmeier, C., 1993. Internal atmospheric gravity

waves near the coast of Antarctica, Boundary-Layer Meteorol., 66, 1–17.

Emter, D. & Zurn, W., 1985. Observations of local elastic effects on Earth

tide tilts and strains, in Earth Tides, Benchmark Papers in Geology Series,

pp. 309–327, ed. Harrison, J.C., Van Nostrand Reinhold, New York.

Exß, J. & Zurn, W., 2002. Reduction of noise in horizontal long period seis-

mograms using local atmospheric pressure, Bull. Inf. Marees Terrestres,137, 10 887–10 892.

Farrell, W.E., 1972. Deformation of the earth by surface loads, Rev. Geophys.Space Phys., 10, 761–797.

Ferreira, A.M.G., d’Oreye, N., Woodhouse, J.H. & Zurn, W., 2006. Com-

parison of tilt data with long period seismograms: surface waves and

Earth’s free oscillations, J. geophys. Res., 111, B11307, doi:10.1029/

2006JB004311.

Forbriger, T., 2007. Reducing magnetic field induced noise in broad-band

seismic recordings. Geophys. J. Int., 169, 240–258, doi:10.1111/j.1365-

246X.2006.03295.x

Freybourger, M., Hinderer, J. & Trampert, J., 1997. Comparative study of

superconducting gravimeters and broad-band seismometers STS-1/Z in

seismic and subseismic frequency bands, Phys. Earth planet. Inter., 101,203–217.

Gossard, E.E. & Hooke, W.H., 1975, Waves in the Atmosphere, Elsevier,

Amsterdam, 456 pp.

Große-Brauckmann, W., 1979. Untersuchungen der Bodenunruhe im Perio-

denbereich von 2 min bis 1 h mit dem Askania-Bohrlochneigungsmesser,

Dissertation, Techn. Univ. Clausthal, 130 pp.

Harrison, J.C., 1985. Earth Tides, Benchmark Papers in Geology Series, Van

Nostrand Reinhold, New York, 419 pp.

Hines, C.O., 1972. Gravity waves in the atmosphere, Nature, 239, 73–78.

Jones, R.M., 2005. A general dispersion relation for internal gravity waves

in the atmosphere or ocean, including baroclinicity, vorticity, and rate of

strain, J. geophys. Res., 110, D22106, doi:10.1029/2004JB005654.

King, G.C.P., Zurn, W., Evans, R. & Emter, D., 1976. Site corrections for

long-period seismometers, tilt- and strainmeters, Geophys. J. R. astr. Soc.,44, 405–411.

Kroner, C., Jahr, T., Kuhlmann, S. & Fischer, K.D., 2005. Pressure-induced

noise on horizontal seismometer and strainmeter records evaluated by fi-

nite element modeling, Geophys. J. Int, 161, 167–178, doi:10.1111/j.1365-

246X.2005.02576.x.

Kurrle, D. & Widmer-Schnidrig, R., 2006. Spatiotemporal features of the

Earth’s background oscillations observed in central Europe, Geophys. Res.Lett., 29, L24304, doi:10.1029/2006GL028429.

Lambotte, S., Rivera, L. & Hinderer, J., 2006. Vertical and horizontal seis-

mometric observations of tides, J. Geodyn., 41, 39–58.

Mockli, A., 1988. Versuche zur Luftdruckabschirmung langperiodischer

Seismometer, Diploma thesis, Geophysics, ETH Zurich, Switzerland,

93 pp.

Muller, T. & Zurn, W., 1983. Observation of gravity changes during the

passage of cold fronts, J. Geophys., 53, 155–162.

Nappo, C.J., 2002. An Introduction to Atmospheric Gravity Waves, Int. Geo-phys. Ser., 85, Academic Press, Amsterdam, 276 pp.

Nawa, K., Suda, N., Fukao, Y., Sato, T., Aoyama, Y. & Shibuya, K., 1998.

Incessant excitation of the Earth’s free oscillations, Earth Planets Space,50, 3–8.

Neumann, U., 1997. Langperiodisches seismisches Rauschen durch atmo-

spharische Wellen, Diploma thesis, Geophysics, University of Karlsruhe,

137 pp.

Neumann, U. & Zurn, W., 1999. Gravity signals from atmospheric waves

and their modeling, Bull. Inf. Marees Terrestres, 131, 10 139–10 152.

Nishida, K. et al., 2005. Array observation of background atmospheric waves

in the seismic band from 1 mHz to 0.5 Hz, Geophys. J. Int., 162, 824–840,

doi:10.1111/j.1365-246X.2005.02677.x

Olsen, N., 1998. The electrical conductivity of the mantle beneath Europe

derived from C-responses from 3 to 720 hr, Geophys. J. Int., 133, 298–308.

Riccardi, U., Hinderer, J. & Boy, J.-P., 2007. On the efficiency of barometric

arrays to improve the reduction of atmospheric effects on gravity, Phys.Earth planet. Inter., 161, 224–242.

Richter, B., Wenzel, H.-G., Zurn, W. & Klopping, F., 1995. From Chandler

wobble to free oscillations: comparison of cryogenic gravimeters and other

instruments in a wide period range. Phys. Earth planet. Inter., 91, 131–

148.

Rodgers, P.W., 1968. The response of the horizontal pendulum seismometer

to Rayleigh and Love waves, tilt, and free oscillations of the earth, Bull.seism. Soc. Am., 58, 1384–1406.

Sorrells, G.G., 1971. A preliminary investigation into the relationship be-

tween long-period seismic noise and local fluctuations in the atmospheric

pressure field, Geophys. J. R. astr. Soc., 26, 71–82.

Steffen, H., 2005. The importance of instrument location on barometric

pressure-induced noise, Bull. Inf. Marees Terrestres, 141, 11 293–11 302.

Steffen, H., Kuhlmann, S., Jahr, T. & Kroner, C., 2005. Numerical modelling

of the barometric pressure-induced noise in horizontal components for the

observatories Moxa and Schiltach, J. Geodyn., 41, 242–252.

Tsai, V.C., Kanamori, H. & Artru, J., 2004. The morning glory wave

of southern California, J. geophys. Res., 109, B02307, doi:10.1029/

2003JB002596.

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Page 17: On reduction of longperiod horizontal seismic noise …geo.mff.cuni.cz/~jz/tmp/presentations/chile2010/zurn...On reduction of long-period horizontal seismic noise using local barometric

796 W. Zurn et al.

Webb, S.C., 2002. Seismic Noise on Land and on the Sea Floor, in Int. Hand-book on Earthquake and Engineering Seismology, Part A, Chapter 19,

pp. 305–318, eds Lee, W.H.K., Kanamori, H., Jennings, P.C. & Kisslinger,

C., Academic Press, Amsterdam.

Wielandt, E., 2002. Seismometry. Chapter 18 in Int. Handbook on Earth-quake and Engineering Seismology, Part A, pp. 283–304, eds Lee, W.H.K.,

Kanamori, H., Jennngs, P.C. & Kisslinger, C., Academic Press, Amster-

dam.

Wielandt, E. & Streckeisen, G., 1982. The leaf-spring seismometer: design

and performance, Bull. seism. Soc. Am., 72A, 2349–2368.

Yuan, X., Kind, R. & Pedersen, H.A., 2005. Seismic monitoring

of the Indian Ocean Tsunami, Geophys. Res. Lett., 32, L15308,

doi:10.1029/2005GL023464.

Zurn, W. & Neumann, U., 2002. Simplistic models of atmospheric effects in

horizontal seismograms, Bull. Inf. Marees Terrestres, 137, 10 875–10 879.

Zurn, W. & Widmer, R., 1995. On noise reduction in vertical seismic records

below 2 mHz using local barometric pressure, Geophys. Res. Lett., 22,3537–3540.

Zurn, W. & Widmer, R., 1996. World-Wide observation of bichromatic long-

period Rayleigh-waves excited during the June 15, 1991 eruption of Mount

Pinatubo, in Fire and Mud: Eruptions and Lahars of Mount Pinatubo,Philippines, pp. 615–624, eds Newhall, C. & Punongbayan, R., Philippine

Inst. of Volcanol. Seismol., Quezon City and University of Washington

Press, Seattle.

Zurn, W. & Wielandt, E., 2007. On the minimum of vertical seismic

noise near 3 mHz, Geophys. J. Int., 168, 647–658, doi:10.1111/j.1365-

246X.2006.03189.x

Zurn, W., Emter, D. & Otto, H., 1991. Ultra-short strainmeters: Tides are in

the smallest cracks, Bull. Inf. Marees Terrestres, 109, 7912–7921.

Zurn, W., Laske, G., Widmer-Schnidrig, R. & Gilbert, F., 2000. Observation

of Coriolis coupled modes below 1 mHz, Geophys. J. Int., 143, 113–

118.

Zurn, W., Widmer-Schnidrig, R. & Bourguignon, S., 1999. Efficiency of

air pressure corrections in the BFO-Records of the Balleny Islands

Earthquake, March 25, 1998, Bull. Inf. Marees Terrestres, 131, 10 183–

10 194.

C© 2007 The Authors, GJI, 171, 780–796

Journal compilation C© 2007 RAS

Administrator
Highlight
Administrator
Highlight