Sie sind auf Seite 1von 14

Geophysical Prospecting, 2001, 49, 431444

Porosity and permeability prediction from wireline logs using artificial neural networks: a North Sea case study
Hans B. Helle,1* Alpana Bhatt1,2 and Bjrn Ursin2
1

Hydro E&P Research Centre Bergen, Sandsliveien 90, 5049 Sandsli, Norway, and 2Norwegian University of Science and Technology, Department of Petroleum Engineering and Applied Geophysics, 7491 Trondheim, Norway Received June 2000, revision accepted March 2001

ABSTRACT Estimations of porosity and permeability from well logs are important yet difficult tasks encountered in geophysical formation evaluation and reservoir engineering. Motivated by recent results of artificial neural network (ANN) modelling offshore eastern Canada, we have developed neural nets for converting well logs in the North Sea to porosity and permeability. We use two separate back-propagation ANNs (BPANNs) to model porosity and permeability. The porosity ANN is a simple threelayer network using sonic, density and resistivity logs for input. The permeability ANN is slightly more complex with four inputs (density, gamma ray, neutron porosity and sonic) and more neurons in the hidden layer to account for the increased complexity in the relationships. The networks, initially developed for basin-scale problems, perform sufficiently accurately to meet normal requirements in reservoir engineering when applied to Jurassic reservoirs in the Viking Graben area. The mean difference between the predicted porosity and helium porosity from core plugs is less than 0.01 fractional units. For the permeability network a mean difference of approximately 400 mD is mainly due to minor core-log depth mismatch in the heterogeneous parts of the reservoir and lack of adequate overburden corrections to the core permeability. A major advantage is that no a priori knowledge of the rock material and pore fluids is required. Real-time conversion based on measurements while drilling (MWD) is thus an obvious application.

INTRODUCTION Porosity and permeability are the key variables in characterizing a reservoir and in determining flow patterns in order to optimize the production of a field. Reliable predictions of porosity and permeability are also crucial for evaluating hydrocarbon accumulations in a basin-scale fluid-migration analysis and to map potential pressure seals in order to reduce drilling hazards. Several relationships have been offered which can relate porosity to wireline readings, such as the sonic transit time

Paper presented at the 61st EAGE Conference Geophysical Division, Helsinki, Finland, June 1999. *E-mail: hans.b.belle@nho.hydro.com

and density logs. However, the conversion from density and transit time to equivalent porosity values is not trivial. The common conversion formulae contain terms and factors that depend on the individual location and lithology, e.g. clay content, pore-fluid type, grain density and grain transit time for the conversion from density and sonic logs, respectively, that in general are unknowns and thus must be determined from rock sample analysis. Permeability is also recognized as a complex function of several interrelated factors such as lithology, pore-fluid composition and porosity. Thus, permeability estimates from well logs often rely upon porosity, e.g. through the KozenyCarman equation, which also contains adjustable factors such as the Kozeny constant, which varies within the range 5100 depending on the reservoir rock and grain

q 2001 European Association of Geoscientists & Engineers

431

432 H.B. Helle et alX

geometry (Rose and Bruce 1949). Nelson (1994) has given a detailed review of these problems. Motivated by recent results of artificial neural network (ANN) modelling by Huang et al. (1996) and Huang and Williamson (1997) applied to porosity and permeability prediction from offshore Canada wireline data, we applied a similar approach to data from the North Sea. The objective here is thus not to propose a new method, but to develop networks that are applicable to the area at hand. In this case study we have developed networks using data from wells in the Viking Graben. The networks were initially developed for basin-scale pressure analysis in the northern Viking Graben, which is an active area of exploration and production with several producing fields implying access to core and wireline data relevant for a basin-scale flow study. While testing our networks on available core data from hydrocarbon reservoirs in the area, we realized that the prediction accuracy was sufficient to meet normal requirements in reservoir engineering. Thus, to improve the capability of the network to account for variations in reservoir fluid, we added a few training facts from the gas- and oil-bearing sections to the original training data set. We find that our modified general Viking Graben neural nets display a better performance for the main reservoir interval of an oilfield than the specialized networks based only on local core data. THE LOG CONVERSION PROBLEM Geophysical well logs generally provide a better representation of in situ conditions in a lithological unit than laboratory measurements because they sample a larger volume of rock around the well and provide a continuous record. However, as with most well-logging measurements, the sonic log does not directly measure the parameter with which it has become associated, i.e. the porosity, given by

the compressional-wave velocity in consolidated siliciclastics, these can be subdivided into a number of major petrophysical groups according to their clay content, and a consistent velocityporosity transform can be established for each group. From ultrasonic measurements on brine-saturated samples, Klimentos (1991) has provided an empirical formula relating the compressional velocity to porosity f (volume fraction) and clay content Cl (volume fraction) by V P in kmas 5X87 2 6X99f 3X33ClX 3

It thus seems obvious that no single log measurement is sufficient to obtain reliable values of porosity. Additional data would be required from the pore fluid and grain material, which normally are not at hand except for special studies in cored reservoir intervals. This is shown in Fig. 1 where we have combined wireline readings of sonic velocity

fDt

Dt g Dt Y Dt g Dt f

where Dtg and Dtf are the sonic transit times of the grain material and pore fluid, respectively, and Dt is the bulk transit time (Wyllie, Gregory and Gardner 1956). Similarly, the porosity from bulk density log values r requires that the grain density r g and fluid density r f be known quantities, i.e.

fr

rg r rg rf

As demonstrated in studies by, for example, Vernik (1997) of

Figure 1 (a) Empirical velocity-to-porosity transform obtained from linear regression by combining logs (sonic, density) and laboratory data (grain density, clay content and total organic carbon) from the northern Viking Graben. The porosity is obtained from the density using equation (2). A change of 50% (vol.) in the clay content Cl has approximately the same effect on VP as a change of 10% (vol.) in the total organic content TOC as indicated by the arrow. (b) By accounting for Cl and TOC the velocityporosity transform may be significantly improved. The result from equation (3) of Klimentos (1991) is shown for comparison.

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

Poreperm prediction by neural nets

433

and laboratory data of grain density r g. A more consistent velocityporosity transform can clearly be obtained if the clay content is taken into account. Moreover, by including the total organic content TOC as the third independent variable we have demonstrated that the linear least-squares fit can be further improved. However, these relationships are of limited practical value because they require the clay content and organic content to be accurately estimated, which is hard to do in practice, given the limitations in the semi-empirical relationships based on the gamma-ray and resistivity logs. The example of an extended laboratory analysis of samples from core plugs and cuttings shown in Fig. 2 demonstrates the sensitivity of the gamma-ray log readings to variations in the total organic content TOC in a North Sea well. A correlation between clay content and the gamma ray is clearly seen. On the other hand, a pronounced peak in the gamma ray coincides with the peak values of the TOC in the Kimmeridge Clay Formation, showing that the gamma ray also has a strong response to petrophysical variables other than the clay content. Again, this demonstrates that a single log cannot by itself resolve a petrophysical property. Alternatively, a suite of different logs in combination may be used to quantify a given petrophysical property provided its relationship to the log readings can be established. Except for the unknown values of grain material and fluid properties, the porosity can be expressed by linear functions of sonic transit time (equation (1)) and bulk density (equation (2)). Because the sonic and density logs respond differently to the fluid and grain material, and since they constitute independent measurements of the same property, a combination of the two may improve the accuracy compared with that of the log-to-porosity transform based on sonic or density alone. Moreover, by adding the resistivity to the suite of logs, the accuracy of the porosity transform may be further improved since resistivity is normally the best indicator for the type of pore fluid. In the following sections we demonstrate that the sonic, density and resistivity combined into an artificial neural network provide accurate porosity estimates for any combination of grain material and pore fluid. While porosity is fairly linearly related to the sonic and density readings, the common permeability transforms indicate non-linear relationships between permeability and the same physical measurements. As predicted from the Kozeny Carman relationship, the permeability can be expressed by kB

Figure 2 A typical set of laboratory data used in this study. To some extent the gamma ray reflects the clay content. Notice, however, the strong response in the gamma ray to high values of TOC in the Kimmeridge Clay Formation (a) and the large range of grain density variations (b) due to the mixture of the light kerogen (r k < 1.4 g/ cm3) and the heavier clay material (r m < 2.77 g/cm3).

grain or pore diameter and t denotes the tortuosity. The additional dependence on the rock texture, the pore shape and its distribution, along with the clay content, indicates that the relationship between log readings and permeability is more complicated than that for porosity and, moreover, that additional physical measurements are required to represent its value. Because permeability of natural sediments is a tensor rather than a scalar property, anisotropy may further complicate the permeability transform in boreholes that are not normal to the bedding. Commonly used empirically derived equations are of the form (Wyllie and Rose 1950), kB

fx Y Sy v

f3 d 2 Y t

where B is a geometrical factor, d denotes a characteristic

which is similar to the KozenyCarman relationship (4), where Sv is the irreducible water saturation and the parameters B, x and y are determined from data, usually

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

434 H.B. Helle et alX

from a log(k) f diagram. In many cases relationships between permeability and porosity may exist. Schlumberger (1989) provided various published forms of (5). In the following sections we show that accurate conversion from well logs to permeability can be obtained by using the neural network alternative rather than the semi-empirical transforms. As is apparent from the above discussion, a more complex network may be required for permeability compared to that of the porosity network.

B A C K - P R O PA G AT I O N N E U R A L N E T W O R K S The back-propagation artificial neural network (BP-ANN) is a relatively new tool in petroleum geoscience, which is gradually being introduced into several practical applications including seismic analysis. It simulates the cognitive process of the human brain and is well suited for solving difficult problems, such as character recognition, which are not amenable to conventional numerical methods (Lawrence 1994; Patterson 1996; Haykin 1999). The ANN functions as a non-linear dynamic system that learns to recognize patterns through training. The network (Fig. 3) has two major components: nodes (or neurons) and connections (which are weighted links between the neurons). Upon exposure to training examples (patterns), the neurons in an ANN compute the activation values and transmit these values to each other in a manner that depends on the learning algorithm being used. The learning process of the BP-ANN involves sending

Figure 3 Architecture of a BP-ANN with four nodes in the input layer, seven nodes in the hidden layer and only one node in the output layer. The symbols W1i,2j and W2k,3l are the weights connecting the input and hidden layers, and the output and hidden layers, respectively. The two networks used in this study have the same architecture but differ in the number of hidden neurons, i.e. seven neurons in the porosity network and 12 in the permeability network.

the input values forward through the network, and then computing the difference between the calculated output and the corresponding desired output from the training data set. This error information is propagated backwards through the ANN and the weights are adjusted. After a number of iterations the training stops when the calculated output values best approximate the desired values. The similarities between BP-ANN and the common geophysical inversion techniques are obvious. The ANN approach has several advantages over conventional statistical and deterministic approaches. The most important one is that it is free from the constraints of a certain function form. Here, we do not consider procedures, rules or formulae, only what kinds of input data the neural network can use to make an association with the desired output. Moreover, in contrast to linear regression models, the ANN approach does not force predicted values to lie near the mean values and thus it preserves the actual variability in the data (Rogers et al. 1995). A detailed comparison by Huang et al. (1996) of permeability prediction by BP-ANN with those of conventional multiple linear regression (MLR) and multiple non-linear regression (MNLR) techniques clearly favours the BP-ANN approach. There are two questions in neural network design that have no precise answer because they are application-dependent: 1 How much data do we need to train the network? 2 What is the best number of hidden neurons to use? In general, the more facts and the fewer hidden neurons there are, the better. There is, however, a subtle relationship between the number of facts and the number of hidden neurons. Too few facts or too many hidden neurons can cause the network to memorize, implying that it performs well during training, but tests poorly and fails to generalize. There are no rigorous rules to guide the choice of the number of hidden layers and the number of neurons in the hidden layers. However, more layers are not better than few, and it is generally known that a network containing few hidden neurons generalizes better than one with many neurons (Lawrence 1994). For instance, if the relationship between input and output is known to be almost linear, we may emulate the linear regression by choosing the number of independent connections, i.e. neuron weights, equal to the number of independent coefficients in the regression equation. Then, a few neurons may be added to the hidden layer in order to account for non-linearity between input and output. On the other hand, the optimal combination can only be achieved by testing and by learning through experience with the data and problems at hand.

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

Poreperm prediction by neural nets

435

Table 1 Selection of facts for the porosity network Well T-1 T-2 F-4 Q-20 Q-22 Q-23 No. of facts 23 42 2 10 1 3 Porosity range (%) 546 2543 4955 2427 23 2426 Comments Grain and bulk density (Lucas 1998) Grain and bulk density (Lucas 1998) Grain and bulk density (Bhatt 1998) From the gas zone. Core helium porosity High resistivity data. Core helium porosity High resistivity data. Core helium porosity

THE POROSITY NETWORK For the porosity network we used the architecture as shown in Fig. 3 but with only three neurons in the input layer, i.e. density, sonic and resistivity. A single hidden layer has seven neurons and the output layer has only one neuron (porosity). The sources of training data for the porosity network are summarized in Table 1. Training facts are dominated by nonreservoir intervals from Tertiary to Jurassic levels. The majority of the porosity values are based on grain density laboratory measurements and bulk densities from wireline data (Lucas 1998). These data were carefully selected to obtain a range of values appropriate for most sediments in the Viking Graben (Bhatt 1998) for use in a basin-scale fluidflow analysis. Tests of this network reveal excellent overall characteristics when applied to the entire geological section (Fig. 4) as well as in the fine details of a water-bearing reservoir (Fig. 5). The main advantage of using porosity derived from the density measurements is the fact that these are the best possible estimates of in situ porosity values since the compressibility of the pure grain material is likely to be small compared with that of the matrix. The grain density in the laboratory is thus not very different from in situ values, and hence the porosity estimates are less prone to pressure corrections than those based on core plugs (Fig. 6a). On the other hand, the comparison made between predictions and core helium porosity reveals striking similarities (Fig. 5), indicating that core and well data may be fairly consistent. For the initial network the pore fluid was assumed to be brine of density 1.03 g/cm3 since no samples were taken from hydrocarbon-bearing sections. In order to adjust the initial network to account for the various pore fluids, we added a few data points taken from hydrocarbon reservoirs. The training patterns cover the porosity range 0.020.55 (Fig. 7a). From a total of 81 facts only 14 facts are taken from the main test area (Q-field). The capability of the resulting modified network to account for different pore

fluids can be appreciated in Figs 8 and 9. Here we compare the porosity predicted by ANN with those predicted by the densityporosity transform (equation (2)) using a constant grain density r g 2.64 g/cm3 and with fluid densities r f of 0.25, 0.75 and 1.03 g/cm3 for gas, oil and brine, respectively. The corresponding porosity transforms f r as shown in Fig. 8 reveal strong sensitivity to the pore-fluid density, with differences of 0.10.15 fractional units between the results of assuming brine- versus gas-filled rock. Because of the strong response to pore fluid in f r the common procedure

Figure 4 (a) A test of porosity prediction by ANN in well S and (b) a cross-plot of measured (from bulk and grain density) versus predicted porosity reveal consistent results. The test data are unknown to the network. Density, resistivity and sonic logs are the inputs.

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

436 H.B. Helle et alX

Figure 5 The porosity network trained for basin-scale prediction (Fig. 4) also performs excellently in details when tested against helium core porosity data from a reservoir.

Figure 6 Relative changes in (a) helium porosity and (b) water permeability as a function of confining pressure for a selection of rock samples from a special core analysis study of well Q-0. Values of helium core porosity and Klinkenberg-corrected air permeability at atmospheric pressure are provided.

implies correction for mixed saturation of pore fluids. An additional correction term for clay content (Schlumberger 1989) and a variable r g are normally included to improve the accuracy of the transform, implying the need for additional core data and input from log interpretation to obtain a measure of shaliness from the gamma-ray log. However, in the neural

net approach, only log data are required once the network has been properly tuned to the area and reservoir at hand. In general, there is a good fit between the porosity predicted by ANN (f ANN) and the corresponding f r . For the water-bearing reservoir (Figs 8a and 9a), the two predictions are practically coincident and comply very well with the helium core porosity within a mean difference
Figure 7 Histograms displaying the distribution of facts for (a) the porosity and (b) the permeability networks used in this study (see Tables 1 and 2). The porosity facts are based on grain density measurements and hence are independent of pressure. Klinkenberg corrections have been applied to the permeability data.

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

Poreperm prediction by neural nets

437

Figure 8 Testing the capability of the porosity network for different reservoir fluids: (a) water, (b) oil and (c) gas. Plots of porosity from density equation (2), with fluid densities 1.03, 0.75 and 0.25 g/cm3, are shown for comparison. A constant grain density of 2.64 g/cm3 has been applied throughout. Core helium porosity data are shown for comparison. The corresponding error distributions are shown in Fig. 9. Test data are unknown to the network. The sonic data to explain the peak in ANN porosity at the top of the reservoir and the relatively high values in the lower section not detected by the density log are shown in (c).

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

438 H.B. Helle et alX

Figure 9 As Fig. 8, with plots of the difference between helium core porosity and the porosity values predicted from the neural network (f core 2 f ANN). The difference (f r 2 f ANN) between the porosity predicted from the density equation (2) and the neural net is shown for comparison. The mean error is approximately 0.01 with a standard deviation of less than 0.02 porosity fractions. Test data are unknown to the network.

(f core 2 f ANN) of approximately 0.01 and a standard deviation of approximately 0.015 based on the 155 core samples from well Q-1. Similar conclusions are valid for the oil-bearing reservoir (Figs 8b and 9b) using 96 core samples from well Q-24. In the gas-bearing reservoir (Figs 8c and 9c), particularly near the shalesand transition at the top of the reservoir interval, there is peak in f ANN, which is not present

in the core or the density porosity. While the density is virtually constant, the sonic reveals distinct peaks at the top and bottom of the reservoir. This feature has been observed in several of the wells in the Q-field and is thus considered to be a real low-velocity event. In most cases observed, the density log responds to this transition and, moreover, core plugs from the zone normally reflect its high porosity. Thus, in this case,

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

Poreperm prediction by neural nets

439

Figure 10 Water permeability for a range of confining pressures versus Klinkenberg-corrected air permeability at atmospheric pressure, based on the special core analysis data on 11 core plugs from well Q-0.

the ANN prediction based on input from three logs compares less favourably with the helium core porosity than that from density alone. THE PERMEABILITY NETWORK While porosity is a scalar quantity, the rock permeability is a tensor owing to the directional alignment of the pore structure of natural sediments. Even in the reservoir rocks at hand, we find that the ratio of in-bedding to normalbedding permeability may be one to two orders of magnitude. Since logging tools are confined to the direction of the drill bore, it is expected that the log readings are affected by anisotropy to various degrees, depending on the drilling angle. The permeability of the core plugs is normally measured at

atmospheric pressure using air, and the Klinkenberg correction is subsequently applied to convert to equivalent fluid permeability. Standard core permeability data thus represent values at the surface while logs are obtained at in situ conditions in the reservoir, where the confining pressures are more than 200 bars. Compression of the rock changes the pore and the pore-throat-size distribution. Changes in the pores may increase the tortuosity and close some of the fluidflow paths. At the surface the permeability of a core sample may be overestimated by a factor of two compared with its in situ value. In the initial study we attempted to overcome the problems of permeability anisotropy by restricting the selection to vertical wells and in-bedding permeability (kh). However, we find that the variations in the permeability anisotropy are confined to a much smaller scale (,0.1 m) than the spatial resolution of the logging tools (,1 m) and thus anisotropy variations appear to have less impact on the log readings than expected. Correction for the pressure effects is a more difficult problem, which cannot be solved within the present industry practice where only a small number of core samples from a field is used for investigating the effect of overburden pressure. Moreover, there is no obvious procedure to convert air-permeability data at atmospheric pressure to fluidpermeability data at in situ conditions. The results of a special core analysis study shown in Figs 6 and 10 may be indicative of the general trend, but they cannot be used to establish a generally valid function to convert airwater permeability at the surface to fluid permeability at depth. While the porosity at low effective pressure may be overestimated by 515% (Fig. 6a), the corresponding permeability data may have errors of 20100% (Fig. 6b) depending on the rock texture and history of the individual sample. For the permeability network we used the same general

Table 2 Selection of facts for the permeability network Well Q-11 P-10 H7-1 H7-2 H10-1 H12-6 H3-1 Q-20 No. of facts 44 140 1 1 1 1 2 5 Porosity range (%) 334 532 8.5 9.7 11.6 1.5 2.5 2427 Permeability range (kh) 34 mD212 D 35 mD21.7 D 5 nD 25 nD 39 nD 6 nD 0.5 nD20.8 nD 1.6 D26.3 D Comments Air permeability. Core plugs Air permeability. Core plugs Water permeability (Krooss et al. 1998) Water permeability (Krooss et al. 1998) Water permeability (Krooss et al. 1998) Water permeability (Krooss et al. 1998) Water permeability (Krooss et al. 1998) From the gas zone. Air permeability. Core plugs

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

440 H.B. Helle et alX

Figure 12 Comparison of (a) porosity and (b) permeability predictions with core data in well Q-2. Calcite-cemented layers are well expressed by low values of porosity and permeability. The reservoir is oil-bearing and the hole deviation is 128.

Figure 11 Comparison of (a) porosity and (b) permeability predictions with core data in well Q-4. Depths of large scatter in the core data coincide with a fine-layering sequence seen in the cores but not properly resolved by the logging tools. The reservoir is oil-bearing. The hole deviation is 018. Fine-layering heterogeneity in the lower part of the reservoir is clearly expressed in the core data and partly expressed by the variability in the sonic (c) which has best spatial resolution (12 ft). The corresponding variability in the density log is less pronounced. Grain density is shown for comparison.

architecture as above, with four input neurons (density, gamma ray, neutron porosity and sonic), 12 neurons in a single hidden layer and a single neuron (permeability) in the output layer. The sources of training data for the permeability network are summarized in Table 2. Most of the training facts are conventional Klinkenberg-corrected air-permeability measurements on core plugs. In order to tune the initial network for basin-scale applications (Bhatt 1998), a few samples of low-permeability shale data taken from the study of Krooss, Burkhardt and Schlomer (1998) were added. While the porosity network is based on samples from both the Tertiary and the Jurassic, all training facts for the permeability network are confined to cored sections from the upper Jurassic. As can be seen from Table 2, the permeability data are dominated by wells outside the test field (Q-field), and the majority of facts (70%) are from a different field in the same area (P-field). By adding the six low-permeability shale points in the range 0.539 nD to the standard core analysis permeability

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

Poreperm prediction by neural nets

441

Figure 13 Comparison of (a) porosity and (b) permeability predictions with core data in well Q-24. The reservoir is oil-bearing and the hole deviation is 54608.

Figure 14 Comparison of (a) porosity and (b) permeability predictions with core data in well Q-20. There is a gas/oil contact at 3198 m. The hole deviation is 38428.

in the range 34 mD212 D (Fig. 7b), we have covered most sediments within the prospective depths in the Viking Graben. Since most of the facts included in the initial network were taken from water- and oil-bearing rocks, we added a few points from a gas-bearing interval (of Q-20) to cover the complete range of reservoir fluids.

E R R O R S I N C O R E D ATA Enforcing the same measurement conditions for laboratory and log data requires core data obtained under simulated reservoir conditions. The industry practice, however, is to use core data measured at ambient conditions to calibrate log data measured in situ. This practice, which is sometimes necessary for financial reasons or because of technical shortcomings, is scientifically unsatisfactory. When core and wireline data are combined to establish the

networks for quantitative prediction of petrophysical quantities such as porosity and permeability, we should keep in mind possible errors in the data. While the wireline tools measure properties in situ at elevated temperature and pressure, the core data are normally obtained in the laboratory at room conditions. In particular, cores collected at great depths are exposed to mechanical deformation and microcracking that significantly increases the surface values of permeability and porosity compared with those in situ. We may also expect significant scatter in the porosity and permeability data since the mechanical impact may differ for individual rock samples due to different composition and sampling history of the core plug. As shown in Fig. 6, the changes with pressure are particularly strong at pressures approaching atmospheric pressure when the microcracks tend to open. The porosity and permeability versus pressure curves are similar for the majority of the core samples while a few are highly offset from the average curve, indicating that the large amount of scatter in surface values of porosity and

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

442 H.B. Helle et alX

Figure 15 Histograms and cross-plots displaying the difference between values obtained from core measurements and the output from the neural nets for well Q-20 (Fig. 14).

permeability may be due to the different pressure effects on the individual core plugs. While the general trend for permeability (Fig. 10) reveals that highly permeable rocks are more prone to pressure effects than less permeable rocks, one of the samples shown in Fig. 6(b) demonstrates the opposite behaviour. A local or generally valid pressure correction formula is thus not easy to establish. On the other hand, since the air-to-water-permeability conversion seems to be a strong function of permeability itself, some of

the scatter observed in the permeability data could obviously be removed by presenting water permeability at reservoir pressure instead of air permeability at atmospheric pressure. From the results in Fig. 10 we find, for example, that an air permeability of 10 D at atmospheric pressure reduces by 40% to a water permeability of 6 D at 200 bars, while for a 100 mD sample the corresponding reduction amounts to only 15% (85 mD). The scatter in the data, however, is too high to accept the corresponding empirical formula for pressure

Table 3 Summary of coreANN comparison of porosity and permeability

f core f ANN
Well Q-1 Q-2 Q-4 Q-22 Q-23 Q-20 Q-24 Mean 0.007 20.011 20.001 20.011 20.001 0.002 20.014

Kcore KANN

s
0.015 0.065 0.027 0.019 0.024 0.018 0.015

Mean 2454 2222 116 1311 79 2578 2318

s (mD)
1038 769 2414 1426 780 2192 816

Hole angle (8) 12 12 12 63 2930 3842 5460

Pore fluid W O O G O G/O O

Comments

1 training fact porosity 3 training facts porosity 10 porosity, 5 permeability

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

Poreperm prediction by neural nets

443

corrections. Thus, to avoid introducing erroneous overburden corrections to the core data we have, in this study, used the raw Klinkenberg permeability supplied by the core laboratory. However, the problem is of significant practical importance and hence should be investigated further. ERRORS DUE TO RESOLUTION AND S PAT I A L S A M P L I N G Worthington (1991) provided a review of the problems encountered when comparing downhole and core measurements. As with any attempt at combining well logs and core data, shifts between recorded well-log depths and sample depths are possible for a number of reasons. While every attempt is made to remove these depth shifts, undetected depth shifts could cause significant errors in porosity and particularly in the permeability predictions. The spatial scale of the well-log measurements is not equivalent to that of the rock sample measurements. Well-log measurements are more spatially averaged than core data. Permeability and porosity measured from cores are representative of only a small rockmass, while a single well-log reading is a composite result of petrophysical properties within a radius of a few centimetres to several metres, depending on which tool is being used. Small-scale heterogeneity between core samples a few centimetres apart may not be resolved at all by well logs. Due to strong heterogeneity in petrophysical properties, and the anisotropic nature of permeability in most natural rocks, it is often difficult to define a characteristic volume that is suitable for numerical calculations. We must keep in mind that a measured value can serve as an estimate of the property over a small interval. Errors in well-log data are caused by poor borehole conditions. Washout, caving, abnormal mudcake, etc. are all capable of adversely affecting well-log responses. NEURAL NET PREDICTIONS AND C O M PA R I S O N W I T H C O R E D ATA While most of the data used for the network design are taken from various wells in an extended area of the northern Viking Graben, we tested the networks on conventional data from an oilfield. The results of the neural network predictions in the cored reservoir intervals are presented below. Results for a selection of wells are displayed in Figs 1115, and a summary of the error analysis for seven wells with various reservoir fluids and hole deviations is given in Table 3.

In general the error distribution fits the normal distribution. Therefore, the mean values and the standard deviations presented here are those of the Gaussian model. Cross-plots (Fig. 15) are less meaningful since the data in the present situation are dominated by samples in the good reservoir section, with only a few values from low-porosity and lowpermeability rocks. For the seven wells listed in Table 3, the average mean porosity difference (f core 2 f ANN) between the core data and the predictions is less than 1% porosity units, with a minimum of 0.1% for well Q-4 (Fig. 11) and maximum of 1.4% for well Q-24 (Fig. 13; see also Figs 8b and 9b). The average standard deviation in porosity is 2.7%, with a minimum of 1.5% in Q-1 and Q-24 and maximum of 6.5% in Q-2 (Fig. 12). For permeability the differences (Kcore 2 KANN) are more significant, with a minimum of 79 mD for Q-23 and a maximum of 1311 mD for well Q-22. The average standard deviation in the permeability of about 1350 mD reflects the large amount of scatter in the core permeability data and the limited spatial resolution of the logging tools as discussed above. In particular, there is significant small-scale heterogeneity in the lower section of the reservoir as can be seen from the scatter in the core data. However, this feature is less apparent in log data, except for the sonic and density logs as shown in Fig. 11, where a marked increase in the amplitude of short-length variations coincides with intervals where core data exhibit maximum scattering. While the main reservoirs of the wells Q-4 and Q-20 (Figs 11 and 14) look fairly homogeneous in terms of log responses, the corresponding intervals of Q-2 and Q-24 (Figs 12 and 13) are clearly interbedded by calcite-cemented layers of detectable thickness, where the predictions reproduce the core data with reasonable accuracy. Thinner beds not detected by the logs are also present in the well as indicated by the presence of core plugs with low values of porosity and permeability. Obviously, these core plugs taken near the bed boundaries contribute most to the errors given in Table 3. CONCLUSION The neural network approach to porosity and permeability conversion has a number of advantages over conventional methods. These include empirical formulae based on linear regression models or the common semi-empirical formulae, such as Wyllie's equation and the density equation for porosity conversion, and the KozenyCarman equation for permeability conversion. The neural net method represents a

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

444 H.B. Helle et alX

pragmatic approach to the classical log conversion problem that over the years has caused problems to generations of geoscientists and petroleum engineers. Instead of searching for complicated interrelationships between geological/ geophysical properties, the neural net approach requires no underlying mathematical model and no assumption of linearity among the variables. The main drawback of the method is the amount of effort required to select a representative collection of training facts, which is common for all models relying on real data, and the time to train and test the network. On the other hand, once established the application of the network requires a minimum of computing time. For the porosity network we find that porosity values from grain density and in situ bulk density data give more consistent results than using standard helium core porosity data. For the permeability network we normally have no other alternatives than air permeability from core plugs and the network will thus inherit the limitations embedded in the method. Our porosity predictions are sufficiently accurate to satisfy most practical needs. Their accuracy is comparable with that obtained from the density equation. The network approach, on the other hand, requires no a priori knowledge of the grain material and pore fluid, and can thus equally well be applied while drilling without prior petrophysical evaluation. In addition, our permeability predictions are sufficiently accurate for most practical purposes, given the limitations due to the spatial resolution of the logging instruments and the expanded range covered by the permeability values. Application to real-time data (MWD) is the obvious extension of this technique.

REFERENCES
Bhatt A. 1998. Porosity, permeability and TOC prediction from well logs using a neural network approach. MSc thesis, NTNU, Trondheim. Haykin S. 1999. Neural Networks: A Comprehensive Foundation. Prentice-Hall, Inc. Huang Z., Shimeld J., Williamson M. and Katsube J. 1996. Permeability prediction with artificial neural network modelling in the Venture gas field, offshore eastern Canada. Geophysics 61, 422436. Huang Z. and Williamson M.A. 1997. Determination of porosity and permeability in reservoir intervals by artificial neural network modelling, offshore eastern Canada. Petroleum Geoscience 3, 245258. Klimentos T. 1991. The effect of porositypermeabilityclay content on the velocity of compressional waves. Geophysics 56, 1930 1939. Krooss B.M., Burkhardt M. and Schlomer S. 1998. Permeability and petrophysical properties of mudrocks from Haltenbanken area offshore Norway. Report 501398, Institut fur Erdol und Organische Geochemie, Julich, Germany. Lawrence J. 1994. Introduction to Neural Networks: Design, Theory and Applications. California Scientific Software Press. Lucas A. 1998. An assessment of linear regression and neural network methods of porosity prediction from well logs. MSc thesis, University of Reading. Nelson P.H. 1994. Permeabilityporosity relationships in sedimentary rocks. Log Analyst 3, 3862. Patterson D.W. 1996. Artificial Neural Networks: Theory and Applications. Prentice-Hall, Inc. Rogers S.J., Chen H.C., Kopaska-Merkel D.C. and Fang J.H. 1995. Predicting permeability from porosity using artificial neural networks. AAPG Bulletin 79, 17861797. Rose W. and Bruce W.A. 1949. Evaluation of capillary character in petroleum reservoir rock. Petroleum Transactions, AIME 186, 127142. Schlumberger 1989. Log Interpretation: Principles and Applications. Schlumberger Educational Services, Houston. Vernik L. 1997. Predicting porosity from acoustic velocities in siliciclastics: a new look. Geophysics 62, 118128. Worthington P.F. 1991. Reservoir characterization at the mesoscopic scale. In: Reservoir Characterization II (eds L.W. Lake et al.), pp. 123165. Academic Press, Inc. Wyllie M.R.J., Gregory A.R. and Gardner L.W. 1956. Elastic wave velocities in heterogeneous and porous media. Geophysics 21, 41 70. Wyllie M.R.J. and Rose W.D. 1950. Some theoretical considerations related to the quantitative evaluation of the physical characteristics of reservoir rock from electrical log data. Journal of Petroleum Technology 189, 105118.

ACKNOWLEDGEMENTS The work leading to the neural networks used in this study was partly supported by the European Union under the project `Detection of overpressure zones by seismic and well data'. We thank S. Hansen, B. Farrelly and J. Okkerman for important technical comments. We are particularly grateful to the two anonymous reviewers for many detailed corrections and comments that improved the clarity of the manuscript.

q 2001 European Association of Geoscientists & Engineers, Geophysical Prospecting, 49, 431444

Das könnte Ihnen auch gefallen