Anda di halaman 1dari 108
DETECTING ABNORMAL GEOPRESSURE USING SEISMIC REFLECTIVITY by SUNG HOON SEO, B.S. ) THESIS Presented to the Faculty of the Graduate School of The University of Texas at Austin in Partial Fulfillment of the Requirements for the Degree of MASTER OF ARTS 3 THE UNIVERSITY OF TEXAS AT AUSTIN ) May, 1978 ACKNOWLEDGMENTS I wish to express my heartfelt thanks to Dr. Wulf Massell who served as my thesis supervisor and provided guidance during its preparation. I also wish to thank my other committee members, Dr. W. M. Rust, Jx., and Dr. L. E. Long for their suggestions. I thank Dr. M. M. Backus and Dr. C. Wilson, who helped me through their discussions and criticism. I am deeply indebted to Dr. D. G. Bebout for the encouragement and assistance he has given to me while T have been working as his research assistant at the Bureau of Economic Geology. I also appreciate the assistance of Dr. R. G. Loucks and Mr. R. A. Gregory of the Bureau of Economic Geology for their editorial assistance. Funds for reproduction and final preparation of this thesis were pro- vided by the U.S. Department of Energy, Division of Geotherm- al Energy, through the Bureau of Economic Geology- This thesis was submitted to the committee in April, 1978. iii DETECTING ABNORMAL GEOPRESSURE USING SEISMIC REFLECTIVITY ABSTRACT Simple mathematical models of the seismic velocity profile expected in a transition zone from normal to over- pressured rocks predict that the zones should behave like lowpass frequency filters to plane seismic compressional waves. This phenomenon can be observed on good quality, broad-band seismic sections having a high signal-to-noise ratio. The amplitude and frequency spectra of model re- flections from the transition zone itself can be shown to depend on the magnitude of the geopressure zone and on the thicknesses of the transition zone. Several possible transitional velocity profiles are presented here using observed relations between compression- al wave velocity, pore fluid pressure and density. Gulf Coast well data confirmed the velocity profile derived from fluid pressure in the normal zone and overpressured zone The thickness of the transition zone and the magnitude of the geopressure itself are varied to illustrate their effect on the reflection impulse response. According to amplitude analysis of reflected waves from model transition zones, a geopressure gradient less than 0.7 psi/ft and a iv thickness of transition zone greater than 500 feet cannot be distinguished from normal lithologic boundaries in re- flection seismograms. A synthetic plane wave reflection seismogram was computed for a well that penetrated into a geopressured zone. ‘This was studied in light of the computed model results. Finally a sixteen kilometer seismic reflection profile from Offshore Area One (Line 15) was processed and analyzed to test the feasibility of using the frequency spectra of reflection events to map the location and thick- ness of geopressured transition zones. With the aid of a velocity analysis performed on Line 15, the bandpass fil- ter proved to be useful to map the geopressured layer on the seismic reflection section of Line 15. TABLE OF CONTENTS INTRODUCTION. 2 eee ee ee ee GEOPRESSURE - ee eee ee eee ee General Concept of Geopressure. .... - Mechanisms of Geopressure ...-- ++ - A GEOPRESSURE MODEL... 2 ee ee ee ee ee A Geologic Model.» 2... 1 ee ee ee Hydrodynamic Analysis... s+ ee ee Pressure-Velocity Relationship |... - REFLECTIVITY OF A LINEAR TRANSITION ZONE. . « Literature Review. 2... + +e eee Analytical Solutions. ... +--+ 4+ Approximate Solutions... ..... Reflection Characters of Linear Transition Zone-Model Study... . er Iterative Matrix and its Application to Synthetic Seismograms ....- +. Error Analysis of the Approximate Method. SYNTHETIC SEISMOGRAM OF GEOPRESSURED LAYERS. . Reflection Characters of Geopressure. . . Program for Synthetic Seismogram-SYNBR. . Analysis of Synthetic Seismogram. . . . « FIELD DATA ANALYSIS... ee ee ee eee CONCLUSIONS. © ee ee ee ee ee ee APPENDIX 2 1 eee ee ee ee ee ee REFERENCES CITED. . ee eee eee ees VITA Se ee ee vi Page 1 ay ly 33 40 40 43 48 51 57 66 66 68 70 82 93 95 96 100 Figure 10 a 12 13 4 LIST OF FIGURES Page Formation pressure versus depth en- countered in the five wells along the Texas Gulf Coast... - ee ee eee ee eS Top of geopressured zone, South Texas... . 6 Well log responses in the transition zone. . & Plot of interval transit time versus depth for sands and shales. .-..-- + + 10 Relationship between depth, pressure and hydraulic conductivity (k) for Gulf Coast ee ee ee 13 Dehydration zones of clay minerals of South Texas. se ee ee ee ee ld Marine shale bulk composition during dehydration... ee ee eee ee ee 16 Geologic model of a geopressured shale mass under a marine depositional environment. . - 19 One-dimensional thermodynamic model. . . . - 23 Qualitative comparison of heat flow solution with pressure distribution and geopressure analogy. settee ee es Distribution of geopressure in the transi- tion zone at different pressure gradients. . Product of hydraulic conductivity (K) and time (t) versus thickness of transition zone. se ee eee ee Distributions of geopressure in the transition zone at different thicknesses of transition zone... ee ee ee Relationship between hydraulic conductivity (K) and time (t) with the thickness of transition zone as a variable parameter. . 27 29 30 31 32 vii Figure 15 16 a7 18 1g 20 al 22 23 24 25 26 velocity of a water saturated sand- stone core as a function of skeleton pressure . Computed velocity profiles of geo- pressure model I-a, I-b, I-c and I-d using equation 7..---.-+--- Computed velocity profiles of geo- pressure model II-a, II-b, II-c and TI-d using equation Tee eeereae Interval transit time of shales in the Socony Mobil #B-1 Chapman Well is compared to theoretical shale compaction trend calculated by equation 7... . - Ramp transition model... 2. eee Reflection coefficients in terms of frequency ratio to cutoff frequency (f calculated by the Wolf's (1937) analyti- cal solution for a linear transition model... -- +s Width of boxcar in time domain con- trols frequency band... + 1 es Schematic illustrations of velocity functions of a linear transition zone in space domain and in velocity increments depth increment and (b), and reflection pair (c)h. ee ee time domain (a), of the pair by unit unit time increment coefficients of the Two groups of ramp transitions models ...--- Amplitude spectrums of impulse responses of ramp transition models... ...-- Waves incident, reflected and transmitted froma stack of layers... ... ++ Comparisons of reflectivity of discrete model with analytical solution at dif- ferent number of layers (or number of samplings)... «+ viii Page 35 37 38 39 42 52 53 55 56 58 64 Figure 27 28 29 30 31 32 33 34 35 36 37 38 39 Error analysis relational to number of layers (or number of samplings) in a transition zone... ee ee Impulse responses of geopressured zones as shown by Geopressure Model Group Ina, I-b, Inc, rd... ee ee ee Impulse responses of geopressured zones as shown by Geopressure Model Group II-a, II-b, II-c, and TI-d. 2 eee ee ee Reflectivity functions. .... 1... eee A frequency analysis of reflection wavelet in Offshore Area 1, Line 15...-.+.- Location map of seismic survey lines and wells of Offshore Areal... . + e+e ee Velocity analysis near well "A" of Offshore Areal, eet eee ee te eee RMS velocity plot of well "A" at Line 15 of Offshore Areal... 1. ee ee eee Seismic section of Line 15, Offshore Area 1 (normal display)... .. +--+ = Seismic section of Line 15, Offshore Area 1 (filtered by 1-10 Herz bandpass filter)... .-- ee eee eee Seismic section of Line 15, Offshore Area 1 (filtered by 11-20 Herz bandpass filter)... ++. ee ee eee Seismic section of Line 15, Offshore Area 1 (filtered by 21-30 Herz bandpass filter)... 1... e+e ee eee Seismic section of Line 15, Offshore Area 1 (filtered by 31-40 Herz bandpass filter)... .. ee eee eee ix Page 65 73 74 75 7 85 86 87 8s 89 90 91 92 LIST OF TABLES Table Page 1 Model I-a, b, c, de ee eee eee ees 78 2 Model II-a, by) c,d eee eee ee ees 78 3 A South Texas Well. 2 ee ee ee ee es 78 4 Sand/Shale Lithologic Boundary Tertiary Clastic Section (10,000 feet)... ..- 79 INTRODUCTION The mechanism of geopressure build-up has been studied by many authors such as Dickinson (1953), Brede~ hoeft et al. (1968), Burst (1969), Chapman (1972), and Jones et al. (1975) due to the growing activity in petro- leum exploration. The main reasons that geopressure attracts our concern come from (a) the effort to avoid the drilling hazards by predicting the presence of abnormally high pressured zones, and (b) the possibility of using geo- pressured zones as an energy resource Common ways of predicting abnormally pressured zones are drilling experiences in the region, existing well logs of the regions, interpretation of geologic data and geophysical exploration. Velocity analysis, as applied by Pennebaker (1968), has become a commonly used geophysi- cal method of mapping overpressured formations. A direct mapping method using reflection seismograms would be most desirable; however, not a single literature reference is available describing such an approach! This study is aimed at examining one feasible way for detecting geopressures using seismic reflection data. General concepts of geopressure and its geologic and hydrodynamic processes are discussed in the first part 1 of this thesis. This will lead to a discussion of pore pressure-velocity and finally to the reflection charac- teristics of the transition zone. These considerations are then used to construct theoretical reflection seismograms. The observed low pass frequency characteristics of the transition zone is finally examined by special processing of a 24-fold offshore CDP reflection seismogram. GEOPRESSURE General Concept of Geopressure Geopressure in a geologic formation is defined as pore fluid pressure whose magnitude is greater than that of local hydrostatic pressure. Geopressure, therefore ranges between hydrostatic pressure and lithostatic pres~ sure, which is simply the overburden pressure. The latter is the vertical stress due to the weight of both solid and fluid material above. An average Gulf Coast gradient of overburden pressure is taken as 1 psi per foot of depth. Small departures from this value have been noted (Dickin- son, 1953). Hydrostatic pressure is defined as the pressure exerted by a column of free fluid. The hydrostatic gra dient in the Gulf Coast is taken to be 0.465 psi per foot and is based on the density of a solution containing 80,000 ppm sodium chloride (Dickinson, 1953) Abnormally high pore pressure in a sedimentary ba- sin occurs upon sealing of the formation so that pore fluids cannot be expelled by further compaction. As a for~ mation is buried under increasing thickness of overburden, compacting processes tend to reduce the volume of pore space in the formation. If the fluids are free to escape, pore pressure remains normal. When the reservoir is sealed, however, the overburdne is partially supported by 3 interstitial fluids, and further compaction stops. The formation is then said to be abnormally pressured or geo- pressured. According to drilling records from the Texas Gulf Coast, it is not unusual to observe a geopressure gradient of 0.9 psi per foot of depth. A plot of formation pres- sure and pressure versus depth for five Gulf Coast studies, conducted by Bebout et al. (1975), concluded that whereas the depth to the top of the geopressure zone varies re- gionally, it is generally encountered at about 10,000 feet and becomes shallower away from the coast with an average gradient of 80 ft. per mile. A regional contour map of the top of geopressure is shown in Figure 2 (Bebout et al. 1975). It is necessary that a geopressured formation be sealed by an impermeable layer, or at least leak at a slower rate than required to maintain normal pressure. The most common sealing material in the Gulf Coast is shale. It is, therefore, probable that most geopressured formations are the shale masses themselves. Sand layers occurring inside the overpressured shale mass can then become geopressured reservoirs The thickness of the transition zone, from the normally pressured to the abnormally pressured condition, ranges froma few feet to a few thousand feet. The thick- FORMATION P! o 2 4 6 8 0 12 14 SURE, 1,90 7 nnn feet Figure 1. Formation pressure versus depth en- countered in the following wells along the Texas Gulf Coast: ARCO #C-3 Mclean (225-9E-7, Jim Hogg Co.), Socony Mobil #B-1 Chapman (19S-20E-7, Nueces Co.), Shell #A-1 Weatherby (185-12E-6, Duval Co.), Magnolia #2 Wilkinson (128-20E-8, Goliad Co.) and Tenneco #43 McAllen (318-15E-4, Hidalgo Co.). Figure 2. Top of geopressured zone, South Texas. Contoured depths are in feet below sea level (Bebout and others, 1975). ness of the transition zone depends upon the permeability of the sealing rock materials and the length of time since the leakage of fluid started. The pressure gradient of the transition layer depends on the pressure difference between the two formations in contact and on the thickness of the transition layer. Geopressured zones can sometimes be identified on well log records by distinguishable features of the tran- sition zone (Musgrave, 1965). Figures 3a, b, and c are examples showing the transitional features of resistivity, sonic and density logs at the transition zone of a Gulf Coast well. Resistivity values (Figure 3a) decrease from about 1.2 q-m in the normally-pressured region at 10,000 feet to 0.6 a-m in the geopressured zone over a 300-foot interval. This decrease in resistivity is explained by the increase of the amount of pore fluid due to under compac- tion in the pressured formation. Sonic transit time in- creases from 100 ysec/ft in the normal section to 140 H- sec/ft in the geopressured section due to the undercompac- tion of grains. Bulk density also changes from 2.42 to 2.30 grams/om? due to the increase of the amount of pore fluid. A semilogarithmic plot of sonic transit time ver- sus depth is shown in Figure 4 for the same well data shown in Figure 3b. In Figure 4, sand and shale data are plot- ted separately for better resolution of the normal shale ‘sexes, ‘Azunop seoenn ‘TTeM ueudeYys T-d# Ttqow Auos0g Jo ouoz uoTITsuery ay} UT sasuodser BOT TIeM *€ SAnbTa 3NOZ NOILISNVYL [SS (e) @) ca) (®) oos‘ol 4 00r'Ol ZI % Ze = 4 #400701 »» Zs EE % 23 @ Zl g Z Zana, A 4A B Zj00"'! Zz Z B nu 72] 2 Bz Z ae Aoovor | ! L | | 0066 of sz o%@ os co os) Zz 1 0 O HOfub tyfoesrt wey nw ALISNGG OINOS: ALIAIUSISay ds trend. The normal shale-compaction trend falls approx- imately on a straight line on the semi-logarithmic plot. Any departure from the normal trend is interpreted as either a lithologic change or the presence of abnormal pres- sure. In Figure 4 the top of geopressure seems to be lo- cated at about 10,000 feet where the sonic transit time departs from its normal trend. r 100! 20}0 300 I | i oo¢ | { I L 4 eo | SAND TREND; = ® gSHALE TREND I e Lo ! » 8 3 et S g & “rps BL Bi 8 a +10 + 12 Figure 4. Plot of interval transit time versus depth for sands and shales in the Socony Mobil #B-1 Chapman well, Nueces County, Texas. 10 1 Mechanisms of Geopressure A number of authors have suggested various physi- cal and chemical processes that might produce excess fluid pressure in a sedimentary basin. Bredehoeft and Hanshaw (1968) summarize the possible causes of geopressure build- ups: (1) aquifer head, (2) tectonic compression, (3) con- tinuous loading and compaction of sediments (Hubbert and Rubey, 1959), (4) fossil pressure that corresponds to a previous greater depth of burial (Watts, 1948), (5) water derived from magmatic intrusion (Platt, 1962), (6) infil- tration of gas, (7) solution or precipitation of materials with consequent change in pore volume (Levorsen, 1954), 8) phase changes, such as gypsum to annydrite (Heard and Rubey, 1966), montmorillonite to illite conversion (Powers 1967), or breakdown of heavy hydrocarbons into light frac- tions (Chaney, 1949), (9) fluid volume expansion or reduc- tion caused by temperature changes (Levorsen, 1954), and (10) osmotic membrane phenomena (Hanshaw and Zen, 1965). Among these possible mechanisms, continuous load- ing and compaction of sediments and phase changes of mont- morillonite to illite are usually cited to account for the observed geopressures of the Gulf Coast region. Bredehoeft presents an example wherein the fluid pressure of clay layers may approximate lithostatic pressure if the hydrau- lic conductivity K of clay is assumed to be 107° cm/sec 12 (3.3 x 107° ft/sec) and the depositional rate of the basin is 500 m/10® years (or 1500 ft/10° years). These data are similar to those of the Gulf Coast. Figure 5 shows the pressure-depth plot that approximates the conditions of a sedimentary basin receiving sediments at a constant rate as assumed for the Gulf Coast of Louisiana and Texas Burst (1969) presents details of how dehydration of montmorillonite contributes to abnormal pore fluid pres~ sure in South Texas. The stepwise dehydration of clay minerals with increasing depth of burial and with formation temperature is illustrated for one of the Gulf Coast wells in Figure 6. By examining the percent of dehydrated clay with depth, Burst noted three stages of dehydration of clay minerals. At depths less than 8,000 feet (Stage One the percentage of dehydrated clay to mixed clay remains constant at about 15-20 percent. In this zone most of the water is forced out of the pore space of clay by the over- burden pressure. In Stage Two the maximum dehydration of clay takes place in the depth range from 8,000 to 12,700 feet. The percent of dehydrated clay (illite) significant- ly increases from 15-20 percent to 65 percent. The sedi- ment remains in a state of quasi-equilibrium as it absorbs heat transmitted from the deeper geologic section. Accord- ing to Powers (1967) and Burst (1969), clay compaction as a result of interlayer water discharge is postulated to be 13 { sedimentotion = 500 m/10*yr 010% of sedimentot 5, 23% 10%m 10,000 m 5 10 20 2 (PRESSURE HEAD (10? m OF WATER) Figure 5. Relationship between depth, pressure and hydraulic conductivity (kK) for the Gulf Coast assuming sedimentation is at a constant rate, where Ss is specific storage and I is total thick- ness of sedimentation. (Bredehoeft and others, 1968). 14 Figure 6. Dehydration zones of clay minerals of South Texas. (Burst, 1969). 1s a temperature-dependent phase change in which a silicate lattice hydrate (montmorillonite) begins to dehydrate at a critical temperature ranging from 200° to 230°F, regardless of burial depth. Stage Three is, as shown in Figure 6, the restrict- ed dehydration zone at depths below about 12,700 ft. In this stage, the final water layer is gradually forced out of the clay lattice and pore spaces as sediment temperature increases. The rate of continued dehydration during this last stage is apparently slow, even by geologic standards. The volumetric change expected for marine sediments during dehydration is shown in Figure 7. During the three dehydration processes, the original shale mass loses about 93 percent of its water content and shrinks to about one- fourth of its original bulk volume. 16 density=1.22 RECENT BURIAL Ee ES tomtoyer woter Sweling Gly Soe Solide = « ng Non-cley seligs oH cone | 421.96 \ AFTER ‘out a=2.28 DEHYORATION arter m 422.57 ene arter ~~ ~~ __ DeHyORATION Figure 7. Marine shale bulk composition during dehy- dration (Burst, 1969). A GEOPRESSURE MODEL A Geologic Model Consider a very thick, predominantly shale section deposited under marine conditions (Figure 8a) containing clay with minor sand. The horizontal permeability is much greater than the vertical permeability because of the dis tribution of thin sand layers at discrete intervals. Al- though these sand layers may be so thin that they do not produce substantial reflections in the normal seismic re- flection frequency band, they are so permeable that they would release their abnormal pressure immediately if they were brought in contact with normally pressured permeable sandstone layers (Fig. 8a). If a normally pressured per- meable sand layer comes into direct contact with a geo- pressured shale mass above some depth 2=0, then the over- pressured sand layers above z=0 release their excess pres- sure "instantly" to become normally pressured layers (Fig. 8b). The layers below 2=0 start to release their excess interstitial fluid to the overlying layers which have become normally pressured. The flow lines of leaking fluids are all vertical and have no horizontal components because clay layers are assumed to be laterally continuous and homogeneous. Growth faulting is the most common tec- 17 tonic movement in the Gulf Coast region and is considered to cause the phenomena described above. 18 19 “(a Juauercu oTU0}Da7 123 sesso -o0ad otweufpospAy pur ‘(e) uouoAcu OTUOIDEZ eTOJOq JUEUUOZTAUS UOTyTSOd -ep ouTZeu e zopuN sseu eTeYs pernsserdoah e yo Tapow oTboToe) *g aanbTg f Foo ou owen euouzmu 1aAa1 vas LN3W3AOW SINOLI3L 4314 AN3W3AGW SINOLIBL 340438 20 Hydrodynamic Analysis As shown in Figure 8, because flow lines have only vertical components and the pressure gradient of the fluid flow is not linear, this model can be treated as a one- dimensional non-steady state flow problem. The basic equa- tions that govern the flow of fluid in a porous medium have been discussed by a number of authors. Jacob (1950, p. 321- 386) has given a concise development. ‘The governing dif- ferential equation is dh_ Ss 2h az? xk at 0S2<0 @ where h is excess head over normal head at given depth (L), K is hydraulic conductivity (L/T), is specific storage (1/L), z is vertical coordinate (L), t is time (7). Hydraulic conductivity, K, is defined by Hubbert (1940) as Ke xBs (2) where k is permeability (L), p is density of fluid (4/13), g is acceleration of gravity (L/T?), N is viscosity of fluia (M/LT). In soil mechanics, cgs units are universally used, whereas most water-supply studies are carried out with K in gallons 21 per day per square foot (gpd/ft”). K has the dimension of velocity. The U.S. Geological Survey uses the meinzer unit to measure hydraulic conductivity. The meinzer unit is defined as the flow of water in gallons per day through a cross-sectional area of one square foot under a hydraulic gradient of one psi/ft at a temperature of 60°F (Davis, et al., 1966). The specific storage, 8, is defined as the volume of water taken into storage or discharge per unit volume per unit change in head (Jacob, 1940, and Cooper, 1966). The boundary conditions of equation 1 are h (2,0) = H(z) at t = 0 h (0,t) = 0 att =0 h (z,t) = Hz} at t= 0 where H,(Z) is the initial excess fluid pressure at depth z at time t = 0. The solution of this differential equiation 1 is analogous to that of the heat flow problem for a semi- infinite solid with thermal conductivity k and with an initial temperature 7, of the whole body. The top edge of the solid is assumed to be in direct contact with an ideal thermal conductor which was devised to keep the temperature constant at zero degrees (Fig. 9). At any instant of time t = 0, heat begins to flow from the underlying solid to 22 the overlying ideal conductor. It is assumed that the heat flow lines are all vertical because the lateral ex- tent of the solid is infinite. With these initial and boundary conditions, this model reduces to a one-dimension- al thermodynamic problem. The solution for the temperature distribution as a function of Z at any time, t, is Tixt) = 7; ert [ ee] (3) 2ckt where T(z,t) is temperature of the solid at point z and at time t, T, is initial temperature of the solid in degrees centigrade, x is distance measured perpendicularly from the boundary in om, k is thermal conductivity of the solid ex- pressed in cal/sec - cm -°C, t is time passed since the heat started to flow in seconds, and erf is error function defined as erf(x) = Flere. The details of this solu- tion are found in Carlslaw and Jaeger (1959, p. 60). Following the analogy above, the solution for the pressure distribution, H(z,t) can be expressed as H(z,t) = Hy (z,t) (4) and Zz Hg (2,t) = accent] 5 Al 6) where H, is excess fluid pressure at depth point Z and at time t after leakage started, and its unit is given in gm/em? (or 1b/in?), H, |, is initial excess fluid pressure 23 +peqze3s MOTE FeOy Ogze oT] OTFTOedS e ae IT ZoZONPUOD Jo 4D UT oANqeIOdUOI = J pue ‘q903 ut quTod Butazesqo ey3 03 sTeTxezOU Oma UOSMZaq AepUnOG 9U3 wozy e0UR3STP = X oTOYM ‘Tapou OTURUAPOWIEyZ TeUOTSUEUTP-2UD *6 SANBTA oO II YoLoNaNod 0; eanjededway TeTyTUT Seurl MOTs 3veH I ¥ouoadNoo sungereduiay yuezsuoo o 24 (initial fluid pressure less hydrostatic pressure at point Z given in gm/em? (or 1b/in?), 2 is vertical coordinate or distance from the contact surface of the two pressure regimes to the specific point in the overpressured zone given in cm (or 30.487! £t), K is hydraulic conductivity, cm/sec (or 30.487! ft/sec), t is time passed since abnormal fluid pressure started to leak, in seconds, and S, is spe- cific storage of the confining material given in 1/cm (or 30.4871/£t). Geopressure, H(z,t), consists of the sum of hydrostatic pressure and excess pressure, that i H(z,t) = Hy(zptz,t) + HA (2,t) where H, is hydrostatic pressure given as follows: Hy = 0.465+(z,+z), and z, is the depth from the surface to the contact boundary of the two zones given in cm (30.387) ft). Then geopressure can be expressed in the foot-pound- second system as follows: H = 0.465(2,+2) + z + (P,-0.465) ext 6) —_2__| ( loot al where P, is pressure gradient in the geopressured zone g (psi/ft) and 0.0656 comes from the constant term and unit conversion factors, and 0.465 is hydrostatic pressure gra- dient given in psi/ft. The coordinate origin, Z is 0, is put at the con- 25 tact boundary of the two pressure zones. A qualitative comparison of the heat flow solution with pressure distri- bution and the geopressure analogy is shown in Fig. 10. The factors governing the steepness and the thick- ness of the transition zone are the magnitude of the con- fined formation pressure, the hydraulic conductivity of the confining shale layer, K, and the time passed since the leakage started, t. Any additional sedimentation during this time ignored in this model. Bredehoeft et al. (1968), however, considered this effect in the model of Figure 5. Although the specific storage of clay, S, is a factor that influences the curvature and thickness of the transition zone, it is assumed to be constant at 30.48 x 1o-8/et (or 1075/em in c.g.s. unit) as an average value for Gulf Coast clay layers. In applying equation 6 to model this study, the thickness of the transition zone is measured between the top of the geopressured zone and the depth where pressure is within 0.99999 of original pressure. ‘The product of K and t describe the thickness of the transition zone, while initial excess pressure governs the steepness of the tran- sition zone. Figure 11 shows the variation of pressure distributions in the transition zone by changing the geo- pressure gradient from 0.465 to 0.6, 0.8, 0.9 and 1.0 while the thickness of the transition zone is fixed at 300 26 Figure 10, Qualitative comparison of heat flow solution with pressure distribution and geopressure analogy. (a) Temperature distribution in conductor II with distance at time t-0 and t-t,. (b) Pressure analogy at t=0 and t=t,. (c) Geopressure analogy in the formation with depth at time t=0 and t=t,. Here Te = temperature in the transition zone, Ti = initial temperature of conductor II (a), He = excess pressure in the transition zone, Hi = initial pressure of lower medium (b), Hs = hydrostatic pressure, Hi = original geopressure, and He = excess pressure in the transition zone (c). DISTANCE DISTANCE 27 (Q t- tet SMPERATURE T TEMPBRATUR T=" 120 9 _ T=Te jeer x =T) ZONE | ne (b) tet, H PRESSURE Hi 9 t HH; (C) Figure 10 t=t GEOPRESSURE Hi iss TRANSITION ZONE H=Hj +H. 28 feet (Geopressure Model I-a, I-b, I-c, and I-d). Notice that the greater the pressure gradient, Pg becomes, the steeper the slope of the transitional curves becomes. The pressure gradient is usually in the range of 0.465 - 1.0 psi/ft. Figure 12 shows how the hydraulic conductivity, K, and time, t, affect the thickness of the transition zone calculated by equation 6. In this figure it is obvious that the greater the hydraulic conductivity and/or the longer the time since leakage started, the larger the transition zone becomes. As a special case we examine the pressure distribution within a transition zone by changing the thickness of the transition zone, 100, 300, 500 and 1,000 feet respectively when the pressure ratio is fixed at 0.9 psi/ft (Geopressure Model II-a, II-b, II-c, and II- a). This relation is shown on the pressure versus depth plot in Figure 13, where it is noticed that the thicker transition zone has a more gentle slope on the pressure distribution curve. It is obvious from equation 6 and Figures 12 and 13, that hydraulic conductivity and time are principal fac- tors governing the thickness of the transition zone (Figure 14). From this plot we can deduce the thickness of the transition zone by knowing the hydraulic conduc- tivity and the time since leakage began. 29 D-1 seansseid o7qe3soxpAy st 43/tSd S9¥*0 pue ‘(p=1 pue ‘q-1 ‘e-1 [opow eansseadosn) 33/tsd 0°T pue 670 ‘g°0 '9°0 +S3UeTpeXb oanssoad qUerTeJFTP 3e ouOZ UOTITSURIy O43 UT oANSsetdoeh Jo suoTanqT43stTa “TT eanSta 41334 NI Hid3d eeez! eeeet eens 2009 aoer ona 2 I L | @ eae eaar 20H 2009 eves axwovreu eee! UNEIGVED B¥nssaud @aaz! 30 susz0a a ‘euoz uoT}Tsuea} JO ssoUyxoTYI awta pue (x) AgTaTzonpuos oTTNeapAy JO 3onpoza 41334 NI SS3NNOTHL eee! eos ees or oz 45/c01XBv "OE 94 03 paumsse | ‘afeus Jo a#eroys otyToads “S eest eee eese eae "eT eanbra Mae >WIOUHR> 38 *{ wopzenbe bursn p-1I pue O-11 ‘q-1I ‘PIT [epou einsserdoeh yo sattjord AqTOoTea pegndwoy *ZT aanbTa 41334 NI Hid3d eeez! 2200) eevee eeeo ear evo Q 1 | | relqeyaua eT suoe uol4isueay jo sseuyoTUL, 43/Tsd6'0 ST yuatpusd sanssesg aees 2229 aaae eves e005 ae2a! 20011 2002! >WAOUHE> UrNOUL SONIC TRANSIT TI! 50 100 200 psec/ft 1,000 feet DEPTH, So T 12P Figure 18. Interval transit time of shales in the Socony Mobil #B-1 Chapman well is compared to theoretical shale compaction trend calculated by equation 7. 39 REFLECTIVITY OF A LINEAR TRANSITION ZONE Literature Review Reflection coefficient is defined as the amplitude ratio of reflected wave to incident wave. It is well known that the reflection coefficient at a sharp boundary is given as PaYo~ Pay PMot PM, where R is reflection coefficient, p and p, is densities R= (8) of medium 1 and medium 2, V) and V, is velocities of medium 1 and 2. Equation 8 is for a normal incident plane wave. The product pv is the acoustic impedance. A transition zone, the boundary between a normally pressured zone and a geopressured zone, is not as sharp as that between normal sedimentary layers. Velocity and den- sity, or generally, elastic constants, change continuously with depth in the transition zone. A reflected wavelet from a transition zone is expected to be different from that of a sharp boundary both in amplitude spectrum and phase characteristics. Wolf (1937), the first investigator of the reflec- tivity of linear transition zones, defined the term "tran- sition zone" as one in which velocity changes linearly with depth. After Wolf's primary work, the reflectivity of 40 41 transition zones was studied by others using different (1958) and Gupta models. Among these, Berryman et al. (1965) derived analytical solutions for the linear transi- tion models. Peterson et al. (1955) and Sengbush et al. (1961) developed a logarithmic approximate method for mul- tiple transition layers. Wuenschel (1960) also gave an approximate solution using Thomson's (1950) iterative matrix for a linear transition layer. 42 Vo Yo veLocITy + p=1 x=0 L vey(r+ = Vv = i & =H ¥ = 1 Vi Figure 19. Ramp transition model (Wolf, 1937). 43 Analytical Solutions The analytical expression of the reflection co- efficient for a linear transition zone has been given by Wolf (1937). The velocity function of his model is linear with depth and density is assumed to be constant. The reflection coefficient of Wolf's model (Fig. 19) can be obtained by solving a one-dimensional wave equation for a normal incident harmonic plane wave. The wave equation is i: ee ar ox ox where (x,t) is the velocity potential of which derivatives with respect to time give velocity of particles and Vv is the velocity of the P-wave. By solving this differential equation in accordance with boundary conditions at the top and the bottom of the transition zone, the reflection co- efficient in terms of frequency is found as R= 1 (0) 20-2y coth[y Lat] V, is velocity of the top layer of the transition zone. y =V1/4+0% -f=1 44 w= 217E H is thickness of transition zone L is velocity ratio of the bottom layer to the top of transition zone, ‘1 . Yo £ is frequency, and boundary conditions are: a) pressure is continuous across the boundaries b) the vertical component of particle velocity is continuous across the boundary. The details of this deriva- tion are given by Officer (1958, p. 201-207). In general, the reflection coefficient, R, is a complex value. Ampli- tude and phase are dependent on frequency. Figure 20 is a graph of the amplitude of the reflection coefficient as a function of £/f, with the velocity ratio, L, as a variable parameter. Parameter f, is the cut-off frequency whose amplitude is zero in the amplitude spectrum plot, and k is the velocity ratio of the bottom layer to the top layer. The zero amplitudes appear in a series whenever the argu- ment of hyperbolic cotangent, |¥|- 1n L of equation 10 becomes n , where n= 1, 2, 3, - + In Figure 20 one zero is seen on a plot of R versus £/£,. This behavior implies a very important fact that the linear transition layer behaves as a low pass filter for a normal incidence P-wave. a5 Figure 20. Reflection coefficients in terms of frequency ratio to cutoff frequency (fo) calculated by the Wolf's (1937) analytical solution for a linear transition model. Here k = velocity ratio of upper layer to lower layer of transition zone, and density was assumed to be 1.0. 46 Gupta (1965) has derived a more general solution for a liquid medium in which the density and bulk modulus vary according to the laws: pp + bx)” qa) A= Ag (1 + bx)? * (a2) Pis density of the liquid, Ais bulk modulus, pis density of the upper layer or arbitrary constant, Ais bulk modulus of the upper layer or arbitrary constant, and p is arbitrary constant. Gupta's solution, however, has been acquired by a high-frequency approximation. Therefore his solution is applicable only to a model which has a very gradual slope of velocity compared to wavelength. Geopressure transition zones are usually too thin to apply Gupta's solution. Berryman and others (1958) derived an expression of the integrated reflection from a series of ramp transition layers which were simplified expressions of continuous velocity logs. ‘Their solution includes multiple reflec- tions as well as primary reflections from each layer boun- dary. His method gives a very convenient way to construct a synthetic seismogram from a continuous velocity log of a multilayered system. For a single ramp transition zone, Berryman's et al.'s solution is no different from that of Wolf. 47 48 Approximate Solutions Peterson and others (1955) introduced a way to syn- thesize a seismogram from a continuous velocity log. Their idea is based on the fact that a continuously changing vel- ocity function can be approximated by an extremely closely- spaced step function. This approximation was mathematical- ly validated by Bortfeld (1968) for a linear transition layer. The velocity function of a ramp transition layer can be approximated by a series of infinitesimal step func— tions, each standing for a small increment of velocity. Each small step gives rise to a small reflected pulse with appropriate amplitude and polarity where the density is assumed to be constant. The reflected wave energy observed at the surface is the summation of all those small indivi- dual reflected pulses. Peterson's idea starts from the basic equation of the reflection coefficient Pve- OM _ apy P¥2 + BY, 2pv-+4pv (3) where A, is amplitude of the reflected wave, and A; is amp- litude of the incidence wave. If we consider very small increments within the transition layer, then the change in acoustic impedance of this interval will also be very small, so that equation 13 reduces to the approximate expression apy ~ SOY 21, Re apy 2 aln(pv) for R < 0.3 a4) where apv is the incremental change in pv. Also, if the density is related to the velocity in the form of xv", then equation 14 becomes Re talaky™ = [ofan (n+ t)-adny ] o® + z z where k and n are constants. For the Tertiary sand/shale sequences the density- velocity relation has been empirically determined by Gardner and others (1974) to be p= 0.23 + v0-25 (as) This relation leads to a very simple way to synthesize a reflection coefficient sequence that is proportional to the change in the logarithm of velocity. Peterson's equation gives very accurate results where the absolute value of re- flection coefficient is less than 0.3. Sengbush and others (1961) introduced the concept of the reflectivity function defined as a set of reflec- tion coefficients equally spaced in time. By taking a time derivative of the Peterson's equation 14, the reflectivity function r (t) becomes r(t) 4. [sve] (a7) 49 50 where the density is assumed to be constant. Ignoring the constant gain factor, %, the reflectivity function reduces to =a r(t) = [avo] (18) We thus express the reflectivity function as the derivative of the logarithm of the velocity function. A synthetic seismogram of the linear transition zone can be produced by convolving the input wavelet with the reflectivity function along the transition zone. It should be noticed that the synthetic seismogram produced by this method gives only primary reflections without any multiples and transmission losses considered. Wuenschel (1960) developed a different approach to the creation of a synthetic seismogram from a continuous velocity log by making use of Thomson's (1950) iterative matrix. He solved a wave equation by taking Laplace's transformation of variables and applying Thomson's matrix for the multi-layered boundary problem which led to a solution that includes multiple reflections and transmis- sion loss at each boundary. With this method, Wuenschel overcame the weakness of the Peterson method by including multiples and transmission losses. 51 Reflection Characters of a Linear Transition Zone-Model Study For studying the reflection characters of the tran- sition zone, it is essential to examine the reflectivity function. To derive the reflectivity function, the vel- ocity function V(x) should be expressed in terms of a time function V(t). The mathematical derivation of V(t) from V(x) and the derivation of the reflectivity function are given in Appendix A. A linear increase in velocity with depth corres- ponds to an exponential increase in velocity with time. This type of velocity response is called a ramp. Since ramp function is an integrated step function, the reflec- tion from a ramp is an integrated shot pulse of steps. Each pulse has the same amplitude if each pulse is spaced in the equal time interval. This kind of a reflectivity function is often called a boxcar. The Fourier transform of a boxcar is a sinc function which is defined as sin qTt/tt (Bracewell, 1965). Sinc function is a typical low pass filter (Fig. 21). Therefore, the transfer function of a ramp transition zone is a low pass filter. Figure 22 is a schematic illustration of a ramp transition function in space and in time domains (a), velocity increments by unit depth and by unit two-way travel time of the pair (b), and reflection in coefficients of the pair (c). 52 mre Toe TN FRPOUFNGY noMAIN ITS FREQUENCY Hertz Figure 21. Width of boxcar in time domain controls frequency band. (a) CONTINUOUS VELOCITY FUNCTION v V(X) V(t) ‘io DEPTH— TWO WAY TIME— (>) DISCRETE VELOCITY FUNCTION v(t) TWO WAY TIKE— (c) REFLECTION COEFFICIENT Figure 22. of a linear transition zone in space domain domain (a), velocity increments of the pair by unit depth increment and unit time increment (b), and reflection co- efficients of the pair (c). 53 vt Schematic illustrations of velocity function and in time 55 ‘sTapow uorytsuer3 duer so sdnoz6 omy, ‘ez sanbta 388s “ENOZ NOILISNVAL AO SUSSENAOTHL @) (a) (#) 4s as a5 4 dos ok 002 AOL goog | a bessas | I \ g ooo*ét_t oour2t 000" et s (2) _ (a) (2) 308) “ENOZ NOLLISNVAL JO SHSSENWOTHL a as 3g AA ue zg é 3 s 3 . S SS ooo" 4 2 998/43 O00'9L Bs o 8 56 uoq3Tsuex3 duex yo sosucdsea estndur zo sunaz0eds epnzrtduy *sTopow "ye eanbTa zqaoH *AONSHOMYA zqaey ‘sonanosud = | OL os oO oon 0s |S sc\ 5 fs | IS a iS : a Ss iS SWNLOSdS SCQLITAAV 2qa0H *AONENO MA vqaeH ‘KONSNOMHA 2qaeH ‘KONGNOMUE o 004 os ° 004 os ° 001 0s o 18 T | = f lS 6 fel os Sy 3 tt Se |e a a a iB is SWMELOGdS HAN“ I TaN 87 The Iterative Matrix and its Application to Synthetic Seismograms In synthesizing a seismogram of a layered medium Thomson's iterative matrix is one of the basic equations to be applied. The input form of the matrix is a series of reflection coefficients spaced in a unit, sampling time intervals one after another. Where the velocity function is available, the input data set is obtained by taking derivatives of the velocity function. But for a continu- ous velocity log from sedimentary layers or a geopressure transition zone, the velocity functions are not always available. In this case, the most common way to provide input data is to take a sampling from a velocity log with a sampling device or by hand in unit time intervals between each sampling. Iterative matrix theory assumes a stack of layers located between two infinite half-spaces (Fig. 25). The two-way travel time of each layer should be identical. 2 is used as unit time delay operator defined as 2 = eivt, (a9) where w is angular velocity, 27f t is two-way travel time of each layer, equal to the sampling interval, £ is frequency. 58 Figure 25. Waves incident, reflected and transmit- ted from a stack of layers. 59 A normal incidence impulse which has a unit ampli- tude hits the top layer and produces a train of reflected impulses from the boundaries of layers which can be defined by a polynomial as C(z). The transmitted wave is also de~ fined by a polynomial as T(z). C(z) and T(z) are poly- nomials of 2 and are called the z-transformation by intro- ducing the operator z (a) defined by equation 19. The z transform of Thomson'a matrix for the case of Figure 25 is described as: 1 czy] V2" TT F(z) 2*@(4) T(z) (20) Ga) zz) 0 where t is transmission coefficient defined by ty = (1-C;) where t, is transmission coefficient at the ith boundary and C, is the reflection coefficient at the ith boundary, and F(z), G(z),F(1/z), and G(1/z) are all polynomials of z. The general form of F(z) and G(z) are k-1 0 2 F(z) = 1+ £)2 + £227 +... + CCye (21) G(z) 2 - et Zt Ggz + ee + CyeRD (22) where C, is the reflection coefficient of the k th boundary, f£ is the coefficient of variable z in F(z) polynomials, and g is the coefficient of variable z in G(z) polynomials. The polynomials F(z) and G(z) of order k are formed in re- cursive form as in F,(2) = Fyy(2) + Gyez-Gy 5 (2) (23) G2) = Ce Fey (2) + 2G 4 (2) (24) From equation 20 we get the reflected wave and the trans- mitted wave in Z-transform polynomials ci2) = $3 (25) T(z) = fet (26) (refer to Claerbout, 1976, p. 144-157) The iterative matrix has wide applicability to synthesizing the seismograms because it produces both the reflected wave and transmitted wave train. In the reflec- ted wave train it includes multiple reflections as well as primary reflections, and it also takes into account the transmission losses through the boundaries. Also, the matrix does not impose any restrictions on the number of layers or the polarity and magnitude of the reflection co- efficient. For the transition zone, the matrix method is also applicable on the basis that the continuously changing velocity function can be approximated by a series of small step functions. The reflection coefficient of transition zones is usually very small in magnitude because the vel- ecity function is continuous. Therefore, the multiple re-~ 60 61 flections of this zone are negligible. The amplitude of multiples dies out in the third order of primary reflec- tions. Calculations by equation 25 show that if the ampli- tude of primary reflection is in the range between 107? and 107 107 and 107? and so on. , then the first multiple is in the range between 62 Error Analysis of the Approximate Method It is desirable to know what is the minimum number of layers into which the transition zone should be divided to represent a continuous transition zone by a set of dis- crete input data. Model Group I-b of ramp transition zone was examined for the impulse response of each case by dividing the transition zone into 5, 10, 20, 40 layers. For each case, the amplitude spectrum was calculated by equation 25 and Fourier transformation compared with those of an analytical solution by equation 10, from 0 to 100 Hz. A normalized square error (NSE) was calculated by the ) following equation: N Pl. E (a3)... (27) 1 where E” is normalized square error, N is total number of frequency components in the calculation and is total number of input data in the time domain, A, is amplitude of analytical solution of the ath frequency component, and B, is amplitude of matrix solution of the ith fre- quency component that corresponds to (i-1) Hz. In each input signal, the sampling interval was At = 0.5 millisecond, and the total input length (fundamental period 63 of the input signal) was T= .N- at = 256 milliseconds. The amplitude spectrum of the analytical solution and those of the approximate solutions by the iterative matrix method (Claerbout, 1976) are shown in Figure 26. ‘The maximum error in the amplitude spectrum corresponds to the minimum number of layers as shown in Figure 27 where normalized square errors are plotted in terms of the num- ber of layers. NSE decreases logarithmically as the number of layers increases. For the 1l-layer case, the NSE error is about 6.5 x 1074 or 3 percent in average error, and for the 21-layer case, the NSE drops to the order of 2.5 x 10 or 0.5 percent in the average error. According to this error analysis, 21 divisions are sufficient for a transition zone whose thickness is about. 100 feet or about 20 milliseconds in two-way time for the average velocity of 10,000 feet/sec. 64 *(sburtdues yo roqumu ro) sxeAey 30 Joqumu quexezzTp ye UOT|NTOS TeoT ~24yeue Yat Tepour ezer0STp Jo AITATAOeTJaT Jo uosTxeduOD “gz OANBTE ZLY3H) AINANORAA e221 ees ee9 aor a-ez ae ! | eea0a 1 N 3 I ese "2? 4 4 3 a eeai SHHAVI S N \ oO SadaAvT Lb I SUBAVT Le 3 SuBAYI LY eS! 23 4 NOTLN1OS TWOTLITVNY 4 3 ay ee02 a 65 ) Figure 27. Error analysis relational to number of layers (or number of samplings) in a transition zone, SYNTHETIC SEISMOGRAMS OF GEOPRESSURED LAYERS Reflection Characters of Geopressure For mapping georessured layers by the seismic re~ flection method, it is essential that the amplitude of the reflected wavelet from the transition zone of geopressure should be enhanced to the distinguishable level over ambient noise or signals from lithologic boundaries. It is, how- ever, not an easy task to accomplish because the reflected amplitude from transition zone is always weaker than that from the sharp boundary which has the same acoustic impe- dance contrast as integrated impedance contrast of whole transitional boundary. The weakness of the reflected wave from transitional boundary to that from the sharp boundary is due to the attenuation of high frequency components. For the simple linear transition zone, the maximum reflec- ted amplitude can be derived mathematically by using the similarity theorem of the Fourier transform (Bracewell, 1965, p. 101). For the transition zone of geopressure, however, a simple mathematical formula is not as applicable as that of linear transition model, and therefore it should be calculated numerically. These can be done by the synthe- tic seismograms of model geopressures. By acquiring the amplitude spectrum of the reflected wave from geopressured 66 67 zone, we can design the appropriate filter to compensate the weakness of it over that of sharp boundary. Bandpass filters and match filters are considered to be appropriate for this purpose because they allow only the frequency com- ponents in the pass-band, and attenuate those out of the band. Thus, consequently, it can enhance the band-limited signal from the transition zone compared to the signal from a sharp boundary which is assumed to be white. 68 Program for Synthetic Seismogram - SYNBR Although there are a few choices among the programs of synthetic seismograms to model geopressure, SYNBR, the Phoenix System Program, is chosen because this program in- poses fewer restrictions on the input data form, which re- quire the velocity and density data with depths of inter- faces. On the other hand, the iterative matrix program requires reflection coefficients sampled in equi-travel time, which itself needs an additional conversion process to the normal well log data. Also the SYNBR program offers easy application of various band-pass filters of zero phase and minimum phase. The purpose of this program is to create zero off- set synthetic sections, with special applicability to amp- litude anomalies. The input parameters for this program are interval velocity and density above and below the in- terface, depth, dip, and the number of traces to which these parameters apply. From these parameters the program will compute two-way travel time, reflection coefficient transmission coefficient and amplitude of each event (where amplitude - reflection coefficient x accumulated trans- mission coefficient x 32,000 + amplitude of any overlap- ping event). Although the SYNBR program does not produce multi- ple reflections, it does not make any difference because the multiples of the transition zone are negligible in magnitude compared to the primary reflections, as pointed out previously. 69 70 Analysis of a Synthetic Seismogram The main purpose for constructing the synthetic seismograms in this study is to analyze the impulse re- sponses of various models of geopressures and to guide for the designing the best-fitting bandpass filter for the models. Then we can get the resolving power of each model by comparing the amplitudes to one another. As for those of linear transition models, two sets of geopressure models and a well log data are provided for this study. The first group (Geopressure Model I-a, b, c, 4) consists of those models whose pressure ratios are 0.6, 0.8, 0.9 and 1.0 respectively where the thickness of the transition zone remains constant, 300 feet. This group is designed to test the effect of the magnitude of geopressure on the amplitude spectrums of reflected wavelets from the tran- sition zone. The second group (Geopressure Model II-a, b c, @) consists of those models whose thickness of transi- tion zone varies as 100, 300, 500 and 1,000 feet where the pressure ratios remain constant, 0.9 psi/foot. This group is designed to test the effect of the thickness of the transition zone on the amplitude spectrum of the re- flected wave from the transition zone. And the last one is a south Texas well log data which has a geopressured zone at the depth about 10,000 feet. The input data were 7 sampled in ten-foot intervals from 10,000 feet to 12,000 feet. The transition zone appears between 10,000 feet and 10,300 feet. In sampling from the log, Hamming weighting factors (0.23, 0.54, and 0.23) were applied to each sample for filtering out the thin layer effects from the input data. The impulse responses of the models are shown on the right column and their amplitude spectrums are shown on the left column of Figure 28, Figure 29, and Figure 30 corresponding to the Model I set, Model II set and the south Texas well respectively. The maximum amplitudes of Model I increase as the pressure gradient increases, where- as the frequency contents do not change so much because the thicknesses of transition zone are the same for all of this model group, although two-way travel times for the transition zone are not the same among this set. The cut- off frequencies for this group are about 25-40 Herz if the cutoff frequency is defined arbitrarily as the amplitude of any frequency less than 12 decibels below the maximum amplitude. The summary results are shown in Table 1. The maximum amplitudes of Model II set decrease drastically as the thicknesses of the transition zone increase even though the pressure gradients are the same for this set The cutoff frequencies also drastically decrease as the thicknesses of transition zone increase. In other words, 72 high frequency components are severely attenuated, whereas the D.C. amplitudes are not affected. The results are summarized in Table 2. The impulse response of the south Texas well in Figure 30 shows for the depth range from 10,000 feet to 12,000 feet, where the geopressure zone occurs between 10,000 feet and 10,300 feet. ‘The pressure gradient is about 0.9 psi/foot for this well. ‘The amplitude spectrum was taken from the interval of the transition zone. The frequency spectrum of the transition zone of this well shows that the dominant frequencies are 0 to 20 Herz although it looks rather white. The high frequency contents partly come from the big reflection coefficient at the top of the geopressure zone, and somewhat from the irregularity of reflection coefficients in the transition zone. For simulating the natural condition, a band- limited zero phase wavelet (4-40 Herz) was generated and convolved as an input wavelet to models instead of an im- pulse as an input. According to the frequency analysis of OSAI Line 15 data, frequency band ranges are about 0-40 Herz with the dominant frequency band at 4-20 Herz (Fig 31). To distinguish signals from the transition zones over those from the lithologic boundaries, a series of bandpass filters were scanned over each model. A sand/shale litho- logic boundary was simulated and applied to the same series 73 IMPULSE AMPLITUDE SPECTRUM RESPONSE (a) 20 40 60 80 100 120 TINE, 109 msec. PREQUENCY, Hertz (b) TIME, 100 msee. 9 20 40 60 BO 100 120 FREOI CY, Hertz (c) 0 20 40 60 80 100 120 FREQUENCY, Hertz (a) tet 0 20 40 60 80 100 120 TIME, 100 msec. FREQUENCY, Hertz Figure 28. Impulse responses of geopressured zones as shown by Geopressure Model Group I-a, I-b, I-c, I-d. ) IMPUL RESPONSE, (a) 100 msec. (>) 100 msec. (ec) a TIME, 100 msec. (a) se TIMP, 100 msec. Figure 29. shown by Geopressure Model Group II-a, II-b, II-d. AMPLITUDE SPECTRUM 20-4 6080 100 1 FREQUENCY, | 60 6 Hertz ee 20 40 60 FREQUENCY, Her’ 7 20 40 60 80 100 12 FREQUENCY, Hertz reer 0 20 40 60 80 100 12 FREQUENCY, Hertz II-c, al 74 80 100 120 0 0 Impulse responses of geopressured zones as nd 75 REFLECTIVITY FUNCTION(10,000-12,000ft) k, theatre ‘| i FREQUENCY - Figure 30, Reflectivity functions of the Socony Mobil #B-1 Chapman well of South Texas. 76 of bandpass filter to read the effects of bandpass filter for each model. By comparing the maximum amplitudes of the models with that of a sand/shale boundary, we can de- duce the resolving power of the amplitude anomaly for each model, and we can deduce the effects of each bandpass fil- ter for the models. Pass-bands of the filters are 1-10 11-20, 21-30, 31-40 Herz. The outputs of the sand/shale lithologic boundary is also summarized in Table 4. This brine sand/shale lithologic model was simulated for the depth of 10,000 feet in the Tertiary clastic sediment basin. The velocity of brine sand is 10,870 feet/sec and its density is 2.2 gm/cm?; the velocity of shale is 10,300 ft/sec and its density if 2.35 gm/em? (Gardner, Gardner, and Gregory, 1968) According to the analysis of Table 1 above, the maximum amplitudes of the Model I set increase drastically from 109 to 520 as the pressure gradient changes from 0.6 psi/ft to 1.0 psi/ft. The output of the bandpass filter of each model shows that the 1-10 Herz filter reduces the original amplitudes to about 0.4 on average. The 11-20 Herz filter reduces them to about 0.35, the 21-30 Herz to about 0.20 and the 31-40 Herz to about 0.16 of the orig- inal amplitudes. It is evident that 0-20 Herz frequency components are dominant for Model I group. The amplitudes of Model II (Table 2) show that the amplitude decreases 7 0 STRRT TIME 2506 END TIME 3066 DE SPECTRUM oN! NALA i ajo" 20° G0" uo” 0" 80° 7 &0 FREQUENCY - HERTZ (Top of geopressure is about 10,000feet) Figure 31. A frequency analysis of reflection wavelet in Offshore Area 1, Line 15. 78 (2Z*0)/S9E (£7°0) /zb¥ (8T"0)/88Z (S%"0)/STH (07T) /OEST 33 008 60 ZHOb-TE ZHOE-1Z ZHOZ-TT ZHOT-T 2eaqtz-ou Aouenbexzy -uyoTY JueTpeszb (pny {due pezTTeuxou) /epnattdue unutxew gyo-ynd —*suea} eanssead TTOM sexes yanos ve OTaeL (4070) /6L (8T°O)/T2 (6z"O)/Ee (TL*O)/TS_ (0° T) /9TT 8 ooot = P=II (eT°0)/TE (9T*0)/6€ (€£°0)/6L (95"0)/PET (0°T)/8Ez oz 00s 2-11 (pt0)/ss (020) /6L (SEO) /z¥T (6€°0)/SST (0°T)/Z0b oe ooe q-II (2270) /L9% (v2"0)/TOE (L2°0) /zEE (TZ"0)/8S% (0°T)/THET ort oot e-rr ZHOv-TE __2H0€-Tz ZHOZ-1T ZHO! ou Aouenbeay -uxoTUa (pny fidue paztyeuTou) /opnyTTawe _unuTXeu TJO-and csuez3 _Tepow “plo‘q‘e-TI TepoW *Z eTaeL (st°0)/2L (8T"0)/e6 (€€70)/TLT (S70) /09% (0°T) /ozs sz ot P-I (pT70)/S5 (0Z"0) /6L (SE*0) /Z¥T (6€°0)/SST (0°T) /ZOF of 6°0 O-r (vt"0)/ev (0z°0)/T9 (SE*0) /EOT (SE°0)/SOT (0°T) /86z ov 870 a-1 (sto) /9t (2z°0)/¥z (9€°0) /6€ (TE*0)/¥E — (0°T) /60T ov 970 BAT zHov—te __2H0e-12 zHoz-11 2HOI-1 z031}3-0u Aouonbeay yuorpeab (png; idwe poztreuzou) /opnaridue unuwrxeu yyo-3nd__eanssezd _Tepow pfo’q’e-I Tepow *T eTaeL 79 peatoauos (zz0)/9z — (ze70)/8€ (THO) /6b (6¥°0)/8S — (0°T) /6TT eae anduT 2HOb-F peatoauoo (€T°0)/9T — (9T"0) /6T —(BT"O)/TZ (6170) /EZ ~— (OT) /OZT_ FO TEAEA-OW zHOb-TE ZnOe-1z ZHOZ-T1 ZHOT-T___703Tt3-ou TSpny FTdue_pozTTeurou) apna TTdue WRN IeU (3293 000/01) uoTgoeg oT3SeTD AreTaqoL Axepunog oTboTou3tT eTeus/pueS “py STARE 80 drastically from 1241 to 114 as the thickness of transition zone increases from 100 feet to 1,000 feet. The outputs of bandpass filters show that the amplitude of Model II-a which has the shortest transition zone, are attenuated severely by applying the bandpass filters but the attenua- tion ratios are equal for each frequency band. With in- creasing thickness of the transition zone, the attenuation becomes more severe for the high-frequency band than the low frequency band. The south Texas well of Table 3 shows that amplitude attenuations are severe with the application of the bandpass filters, and that 1-10 Herz and 21-30 Herz are dominant frequencies. Now the analysis of Table 4, that is for a sharp lithologic boundary presents an impor- tant reference for assessing the amplitude anomalies of the geopressure models mentioned above. The original amplitude is 120 and attenuation becomes severe by the application of bandpass filters, but the ratios are quite even for each passband because the frequency spectrum is quite white. The ratios are 0.19, 0.18, 0.16, and 0.13 for the bands of 1-10, 11-20, 21-30 and 31-40 Herz respectively. By the convolution with the input wavelet as defined before, however, for the filtered output by each band the low fre- quency components become more abundant. The attenuation ratios are 0.49, 0.41, 0.32, and 0.22 respectively By comparing the original amplitude and bandpassed 81 amplitudes of the sand/shale model to Model I-a, b, c and a, it is concluded that Model I-a has smaller amplitude than the sand/shale boundary, so that a geopressure zone which has the smaller pressure gradient than 0.6, or pos~ sibly 0.7, cannot be detected by the bandpass filter Model II-c has slightly bigger amplitudes than those of the sand/shale boundary except at higher frequency band than 21-30 Herz. ‘The Model II-d has smaller amplitudes than those of sand/shale boundary, therefore 500 feet is the marginal maximum thickness of the transition zone to be detected by the conventional method. The south Texas well model is also detectable by amplitude level, but it seems to be hard to distinguish it from a lithologic boundary because the frequency com- ponents are evenly distributed. FIELD DATA ANALYSIS Line 15 of Off-Shore Area 1 (OSA1) was chosen for the test area to see how the bandpass filters are effective for the prospecting the geopressured horizon. Figure 32 is a location map of survey lines and of wells drilled on the area. According to Well A data, the top of the geo- pressure horizon is located at about 10,000 feet depth. This is also confirmed by the velocity analysis conducted near Well A (CDP 508) where the low RMS velocity zone is detected between about 2.6 seconds and 3.0 seconds of two- way time (Fig. 33). Backus and others (1975) show almost the same result on their RMS velocity plot (Fig. 34). A portion of the seismic section of Line 15 is shown on Figure 35 where the portion between CDP 503 and 511 is displayed in horizontally, and vertically it is shown between 2 seconds and 3 seconds. Figure 35 is normal display of the portion. The outputs of the bandpass filters of 1-10, 11-20, 21-30, and 31-40 Herz are shown in Figures 36, 37, 38 and 39. ‘The horizon at 2.6-2.7 seconds confined in the small square seems to be the top of the geopressure zone, be- cause the depth is almost coincident with the depth of geopressure at Well A which is located in that area, and 82 because the horizon is consistent in normal section and the 1-10 Herz filtered section but rapidly disappears in the high frequency band sections. 83 84 Figure 32. Location map of seismic survey lines and wells of Offshore Area 1 (Backus, 1975). 85 ze eandta ae WaT ols St €V vw a tb 86 L At OF O34 1 AR WEL VELOCITY ANALYSIS W 15 LINE 8 2 Line 15 shows the top of geopressure Velocity analysis near well A of Off- at about 2.6 seconds. shore Area 1, Figure 33. DIP RMS VELOCITY tsicop) | (KFT/SEC) AMP(MV) <4 45 78 9 4 0 4 Toy. 7 | ‘i 1 | ' | | vl i | = BRIGHT SPOT ‘® ROCK o FLAT SPOT X MULTIPLE Figure 34. RMS velocity plot of well "A"at Line 15 of Offshore Area 1. (Modified from Backus, 1975). 87 Figure 35. Seismic section of Line 15, Offshore Area 1 (normal display). Figure 36. Seismic section of Line 15, Offshore Area 1 (filtered by 1-10 Herz bandpass filter). 68 90 OZ%-TT Aq pareatts) T Peay er0yss50 +(s03TT3 ssedpueq zi0H 4g] eUTI JO uoTqDEs OTUSTES "Le eanbra 91 oe-12 Aq pazaaTts) T Perv ex04s330 +(103TTZ ssedpueq 2z0H ‘ST eUTT Jo UoTz005 oTUSTES "Be eanbta 92 Op-Te Aq paxaaTts) T eazy a7048550 + (xeqTTy ssedpueq ‘gt ouPT Jo uoTqDes OTUSTES ‘EL eANbTA 2) 2) 3) 4) CONCLUSIONS Velocity profiles derived from a hydrodynamic model of geopressure fit well with Gulf Coast well log data along normally-pressured zone and abnormally-pressured zone. Error analysis indicates that discrete sampling along transition zone is a valid way to represent the con- tinuously changing transition zone in generating a synthetic seismogram if the number of samples is suf- ficient. Model study of synthetic seismograms shows that: (a the transition zone of geopressure behaves as a lowpass filter; (b) the pressure gradient and the thickness of the transition zone control the amplitude and the fre- quency contents of reflected wave. The bigger the pressure gradient is and/or the thinner the transition zone is, the larger the amplitude and the whiter the spectrum becomes. Amplitude analysis of model geopressure compared with the simulated sand/shale lithologic boundary indicates that: (a) the pressure gradient should be bigger than 0.7 psi/ft, (b) the thickness of the transition zone should be less than 500 feet to give a larger amplitude 93 5) 6) 7 94 than that of a lithologic boundary. The south Texas well study implies that the transitional curve should be smooth to avoid high frequency reflection and to behave as a lowpass filter An analysis of the Offshore Area 1 data suggests that (a) the frequency band at the depth of 10,000 feet is about 0-40 Herz and the dominant frequencies at the top of the geopressure horizon are 4-20 Herz, (b) the band- pass filter is effective under the favorable conditions such as good quality data with big signal/noise ratio because we are dealing with very deep and low-frequency data, and a clean transition zone to distinguish from the lithologic boundary, (c) velocity analysis should be conducted in parallel with a reflection seismogram to give a cross-check of the results. It is left for further study to find means to select boost and sharpen the low-frequency signals in the presence of low-frequency noise and deep lithologic boundaries. APPENDIX I A linear velocity function in terms of depth x is expressed as follows: V(x) vo (2 + Hed x) = vy (1tbx) W where all of the notations were defined in previous sec- tions. Two way travel time of ramp transition layer is 2. [7 ax lo Vix! -2 Veby + Ve Vo b tal Vo | where V(x) = V, (1+bx) e7 ! Therefore x = D By substituting equation 21 into equation 19, velocity in terms of time is obtained: Nebt vit) = vje 2 And reflectivity function is As shown above, the reflection coefficient is function of the slope of the velocity function. 95 REFERENCES CITED Backus, M.M. and Chen, R.L., 1974, Flat Spot Exploration: Presented at 36th European Assoc. of Explor. Geo- physicists, Madrid. Bebout, D.G., Dorfman, and Agagu, 0.K., 1975, Geothermal resources Frio formation, South Texas: Bur. of Econ. Geol., Geol. cir. 75-1. Berryman, L.H., Goupillaud, P.L., and Waters, K.H., 1958, Reflections from multiple transition layers, Part 1 -- Theoretical Results: Geoph. Vol. 23, no. 2, p. 223- 243. Bortfeld, R., 1960, Seismic waves in transition layers: Geophysical Prospecting, Vol. VIII, p. 178-217. Bracewell, R., 1965, The Fourier transform and its applica~ tions: McGraw-Hill Book Co., p. Bredehoeft, J.D. and Hanshaw, B.B., 1968, On the mainte- nance of anomalous fluid pressures -- I. Thick sedi- mentary sequences: Geol. Soc. America Bull., v. 79, p. 1087-1106. Burst, J.F., 1969, Diagenesis of Gulf Coast Clayey sedi- ments and its possible relation to petroleum migra~ tion: Am, Assoc. Petroleum Geologists Bull., v. 53, p. 73-93. carslaw, H.S. and Jaeger, J.C., 1959, Conduction of heat in solids, 2nd edition: London, Oxford Univ. Press, 150 p. Chaney, P.E., 1949, Abnormal pressures and lost circula~ tion, in Drilling and production practice: Am. Petroleum Inst., p. 145-148. Claerbout, J.S., 1976, Fundamentals of geophysical data processing with applications to petroleum prospect- ing: McGraw-Hill Book Co., p. 144-162. Chapman, R.E., 1972, Clays with abnormal interstitial fluid pressures: Am. Assoc. Petroleum Geologists Bull v. 56, no. 4, p. 790-795. 96 97 Cooper, H.H., Jr., 1966, The equation of groundwater flow in fixed and deforming coordinates: Jour. Geophys. Research: v. 71, p. 4785-4790. Davis, S.N., and Dewiest, R.J., 1966, Hydrogeology: John Wiley & Sons, Inc., N.Y., Pp. 156-200. Dickinson, G., 1953, Reservoir pressures in Gulf Coast Louisiana: Am. Assoc. Petrol. Geol. Bull., v. 37, p. 410-432. Gardner, G.H.F, Gardner, L.W., and Gregory, A-R., 1974, Formation velocity and density--the diagnostic basis for stratigraphic traps: Geophysics, vol. 39, no. 6, p. 770-780. Gardner, G.H.F. and Gregory, A.R., 1968, Formation velocity and density-The diagnostics basics of stratigraphic traps: presented at Society of Exploration Geophysi- cists Seminar, Ft. Worth, Texas (March 22, 1968). Gupta, R.N., 1965, Reflection of plane waves from a linear transition layer in liquid media: Geophysics, vol. 30, no. 1, p. 122-132. Hanshaw, B.B., and Zen, E-an, 1965, Osmotic equilibrium and overthrust faulting: Geol. Soc. America Bull., v. 76, p. 1379-1386. Heard, H.C., and Rubey, W.W., 1966, Tectonic implications Of gypsum dehydration: Geol. Soc. America Bull., v. 77, p. 741-760. Hottman, C.E, and Johnson, R.K., 1965, Estimation of for- mation pressures from log-derived shale properties: 3. Pet. Technol., vol. 17, p. 717-723. Hubbert, M.K., 1940, The theory of ground-water motion: Jour. Geol., vol. 48, pt. 1, p. 785-944. Hubbert, M.K., and Rubey, W.W., 1959, Role of fluid pres sure in mechanics of overthrust faulting, I. Mechan- ics of fluid-filled porous solids and its applica~ tion to overthrust faulting: Geol. Soc. America Bull., v. 70, p. 115-166. Jacob, C.E., 1940, On the flow of water in an elastic arte- sian aquifer: Am. Geophys. Union Trans. pt. 2, p. 574- 586. 98 Jones, P.H., 1975, Geothermal and hydrocarbon regimes, northern Gulf of Mexico Basin: Proceedings, First geopressured geothermal energy conference, University of Texas at Austin, June 2-4, p. 15-89. Levorsen, A.I., 1954, Geology of petroleum: San Francisco, W.H. Freeman and Co., 703 p. Musgrave, A.W. and Hicks, W.G., 1965, Outlining shale mas- ses by geophysical methods: Am. Assoc. Petroleum Geologists Bull., Memoir 8, p. 112-136. Officer, C.B., 1958, Introduction to the theory of sound transmission with application to the ocean: McGraw- Hill Book Co., p. 201-207. Pennebaker, E.S., 1968, An engineering interpretation of seismic Presented at the 43rd Annual Fall Meeting of the Soc. of Petrol. Engineers of AIME, in Houston, Texas. Peterson, R.A., Fillippone, W.R., and Coker, F.B., 1955, The synthesis of seismograms from well log data: Geophysics, vol. 20, no. 3, p. 516-538. Powers, M.C., 1967, Fluid-release mechanisms in compacting marine mudrocks and their importance in oil explora tion: Am. Assoc. Petroleum Geologists Bull., vol. 51, no. 7, p. 1240-1254. Reynolds, E.B., May, J.E., and Klaveness, A., 1971, The geophysical aspects of abnormal fluid pressure: Abnormal subsurface pressure, Houston Geol. Soc. Study Group Report, p. 31-47. Sengbush, R.L., Lawrence, P.L., and McDonal, F.J., 1961, Interpretation of synthetic seismograms: Geophysics, vol. 26, no. 2, p. 138-157. Thomson, W.T., 1950, Transmission of elastic waves through a stratified solid material: J. Appl. Phys. 21 p. Watts, B.V., 1948, Some aspects of high pressures in the D-7 zone of the Ventura Avenue field: Am. Inst. Min. Metall. Engineers Trans., v. 174, p. 191-205. Wolf, A., 1937, The reflection of elastic waves from tran- sition layers of variable velocity: Geophysics, v. 2, p. 357-363. 99 Wuenschel, P.D., 1960, Seismogram synthesis including mul- tiples and transmission coefficients: Geophysics, vol. 25, no. 1, p. 106-129. Wyllie, M.R.J., Gregory, A.R., and Gardner, G.H.F., 1958, ‘An experimental investigation of factors affecting elastic wave velocities in porous media: Geophysics, v. 23, no. 3, p. 359. VITA Jung Hoon Seo was born in Sungju-gun, Kyungbuk, Korea on September 14, 1937, the son of Kyungsun Seo and Byungwu Seo. After completing his work at Sungju High School, in 1956, he entered Seoul National University, Seoul, Korea. While an undergraduate, he joined the Korean Army and was discharged after full service. He received the degree of Bachelor of Engineering from Seoul National University in January, 1963. In October of 1962 he was employed by the Geological Survey of Korea as an exploration geophysicist. He participated in a training program of petroleum technology in the U.S.A. sponsored by the Korean Govern~ ment from September, 1972 to May, 1973. He entered the Graduate School of the University of Texas at Austin in January, 1973. Upon returning back to Korea he served as an acting chief of the geophysical section of the Geologi- cal Survey of Korea. In August of 1974, he came back to the Graduate School of the University of Texas at Austin. During his stay as a graduate student he has worked as a research assistant of the Bureau of Economic Geology. Permanent address: 45 Hwajon-dong, Jung-Ku, Daegu, Korea This thesis was typed by Donna L. Precht.

Anda mungkin juga menyukai