Loading...
HomeMy WebLinkAboutCity of Akutan Geothermal Technical Feasibility Report - June 2011 - REF Grant 2195475Akutan Geothermal Resource Assessment Commissioned by City of Akutan, Alaska As part of its Geothermal Development Project June 2011 Principal Investigator: Amanda Kolker, AK Geothermal Other Investigators: Bill Cumming, Cumming Geoscience Pete Stelling, Western Washington University David Rohrs, Rohrs Consulting Akutan Geothermal Resource Assessment 2 Contents Summary ........................................................................................................................................................................ 3 Objectives of Study ........................................................................................................................................................ 3 Introduction ................................................................................................................................................................... 4 Background and Previous Studies ................................................................................................................................. 4 Geologic Setting ........................................................................................................................................................ 5 Geothermics .............................................................................................................................................................. 7 MT Resistivity ............................................................................................................................................................ 7 New Data 2011 .............................................................................................................................................................. 8 1. Temperature Gradient Data ............................................................................................................................. 8 1a. Core Hole Drilling ............................................................................................................................................ 8 1b. End-of-Well Logs ............................................................................................................................................. 9 1c. Equilibrated TG Logs ...................................................................................................................................... 10 1d. P/T Data Analysis .......................................................................................................................................... 13 2. New fluid chemistry and geothermometry .................................................................................................... 15 2a. Sample Collection and Data Sources ............................................................................................................. 15 2b. Chemistry ...................................................................................................................................................... 15 2c. Geothermometry .......................................................................................................................................... 15 2d. Geochemical Model ...................................................................................................................................... 16 3. Core data ........................................................................................................................................................ 16 3a. Overview ....................................................................................................................................................... 16 3b. Rock Types and Primary Mineralogy ............................................................................................................. 17 3c. Secondary Mineralogy, Mineral Paragenesis, and Hydrothermal History .................................................... 17 3d. Permeability and Porosity of Well Rocks ...................................................................................................... 20 Resource Conceptual Models ...................................................................................................................................... 21 Future Drilling Targets ................................................................................................................................................. 25 Capacity Assessment ................................................................................................................................................... 27 Resource Existence and Size ................................................................................................................................... 27 Confidence in Resource Existence ........................................................................................................................... 27 Probable Resource Capacity .................................................................................................................................... 28 An alternative approach .......................................................................................................................................... 28 Monte Carlo Heat-in-Place Option .......................................................................................................................... 28 Resource Risks ............................................................................................................................................................. 29 Upflow Development Risks ..................................................................................................................................... 29 Outflow Development Risks .................................................................................................................................... 30 Conclusions .................................................................................................................................................................. 30 Recommendations ....................................................................................................................................................... 31 References and Bibliography ....................................................................................................................................... 32 Akutan Geothermal Resource Assessment 3 Summary The Akutan geothermal resource can be conceptualized as containing two major zones: an upflow zone and an outflow zone. While the outflow and upflow zones likely represent one interconnected field, they are distinguished here for the purposes of development. The upflow zone temperatures could approach 572 °F (300 °C), and the reservoir probably consists of a brine liquid overlain by a small steam cap. The outflow zone temperatures are lower, decreasing as the brine flows eastward. Fluids produced by corehole TG-2 show evidence of chemical re-equilibration to lower temperatures, with cation geothermometry providing a range from 392-464 °F (200-240 °C). The outflow fluids become extensively mixed with cooler meteoric waters near the surface hot springs. Alteration mineralogy in exploratory coreholes suggests two disappointing conclusions about the outflow system: (1) the rocks in both TG-2 and TG-4 were at temperatures greater than 469 °F (250 °C) in the geological past and have cooled to present temperatures; and (2) the part of the outflow encountered by the wells appears to lack sufficient thickness and permeability to support commercial development. Additionally, development of the shallow outflow would entail significant risk of rapid cooling during exploitation as a result of either cold water influx from near-surface aquifers or injection breakthrough. Exploratory corehole drilling encountered the outflow zone with fluid temperatures of 359 °F (182 °C) at shallow depths of 585’ (187 m). Recent data suggests that the 359 °F (182 °C) zone produced in TG-2 is drawn from a nearby fault zone not located directly below the well. Although it is possible that a hotter resource may exist slightly deeper than either of the current wells, this is unlikely to be the lowest risk target for development. Although TG-2 encountered the outflow predicted near its location, the two exploration coreholes did not demonstrate an outflow resource that would be suitable for development. Given these drilling outcomes and results of new gas geothermometry from the fumaroles, a well targeted to cross the 1500 ft2 (0.5 km2) fumarole field would have the highest probability of encountering commercial production at Akutan. This target is likely to be >428 °F(>220 °C) and could be as hot as 572 °F (300 C). The depth to the target will depend on the elevation of the drill pad but it is likely to be greater than 4000’ (1300m). An important issue is the trade-off between the cost and practicality of constructing a pad closer to the fumarole and drilling further directionally. A 380-428 °F (180-200 °C) outflow resource target about 2200’ (800 m) to the northwest of TG-2 might be preferred if its higher targeting risk and lower generation per well were sufficiently offset by lower drilling and access cost. Objectives of Study This study has three primary objectives: (1) to report on data collection efforts for the Akutan geothermal resource to date; (2) to provide the technical parameters needed when assessing the feasibility of developing the geothermal resource for a combined heat-and-power project envisioned by the City of Akutan and other stakeholders; and (3) to provide well targets for future drilling efforts. This report synthesizes all the datasets collected on the Akutan geothermal field to date (listed on p. 4), and provides an updated assessment of the Akutan geothermal resource based on all available data. Well targets and recommendations for mitigating resource risks are given. Akutan Geothermal Resource Assessment 4 Introduction Akutan Island is located 790 mi (1271 km) southwest of Anchorage and 30 mi (48 km) east of Dutch Harbor. The Island is home to North America’s largest seafood processing plant. The City of Akutan and the fishing industry have a combined peak demand of ~7-8 MWe which is currently supplied by diesel fuel. In 2008, the base cost of power in the City of Akutan was $0.323/kWh (Kolker and Mann, 2009). Since 2008, the City of Akutan has led exploration and other assessment activities in an effort to determine the feasibility of geothermal development on the island. The 2009 exploration program included practical access assessments, a geologic reconnaissance field study, soil and soil gas geochemical surveys, a remote sensing study using satellite data, a review of existing hot springs geochemistry data, a magnetotelluric (MT) survey, and a conceptual model analysis. The 2010 exploratory drilling program included the drilling of slim-hole temperature gradient (TG) wells, fumarole sampling, and chemical analysis of well and fumarole fluids. Follow-up production-size wells are being planned for the near future. Background and Previous Studies As a volcanic island with accessible hot springs, Akutan has been the subject of geothermal resource studies since 1979. The original exploration effort was limited to the immediate hot springs area and included geologic mapping, shallow (<500’ / 150m) geophysical surveys, and fluid geochemical studies. In summer 2009, the COA executed a follow-on exploration program including a geologic reconnaissance field study, soil and soil gas geochemical surveys, a remote sensing study using satellite data, a review of existing hot springs geochemistry data, a magnetotelluric (MT) survey, and a conceptual model analysis. In summer 2010, an exploration drilling program was carried out with two temperature gradient (TG) wells drilled in Hot Springs Bay Valley (HSBV). The following reports have been written on the Akutan geothermal resource, in chronologic order: 1. Motyka, R., and C. Nye, eds., 1988. A geological, geochemical, and geophysical survey of the geothermal resources at Hot Springs Bay Valley, Akutan Island, Alaska. Alaska Division of Geological and Geophysical Surveys (ADGGS), Report of Investigations 88-3. 2. Motyka, R.J., S. Liss, C. Nye, and M. Moorman, 1993. “Geothermal Resources of the Aleutian Arc.” Alaska Division of Geological and Geophysical Surveys (ADGGS) Professional Paper 114. 3. Kolker and Mann, 2009. “Heating up the Economy with Geothermal Energy: A Multi-Component Sustainable Development Project at Akutan, AK.” Transactions, Geothermal Resources Council Annual Meeting 2009. *Both paper and poster format available. 4. Kolker, Cumming, Stelling, Prakash, and Kleinholtz, 2009. “Akutan Geothermal Project: Report on 2009 Exploration Activities.” Unpublished report to City of Akutan and the Alaska Energy Authority, 37p. 5. WesternGeco, 2009. Magnetotelluric Survey at HSBV, Akutan, Alaska: Final Report – 3D Resistivity Inversion Modeling. Unpublished report prepared for the City of Akutan, Alaska, GEOSYSTEM/WesternGeco EM, Milan, Italy, 27p. Akutan Geothermal Resource Assessment 5 6. Kolker, Stelling, and Cumming, 2010. “Akutan Geothermal Project: Preliminary Technical Feasibility Report.” Unpublished report to City of Akutan and the Alaska Energy Authority, 31p. 7. Kolker, Bailey, and Howard, 2010. “ Preliminary Summary of Findings: Akutan Exploratory Drilling Program.” Unpublished report to City of Akutan and the Alaska Energy Authority, 32p. 8. Kolker, Cumming, and Stelling, 2010. Geothermal Exploration at Akutan, AK: Favorable Indications for a High-Enthalpy Hydrothermal Resource Near a Remote Market.” Transactions, Geothermal Resources Council Annual Meeting 2010. *Both paper and poster format available. 9. Rohrs, 2011. “Geochemistry of the Akutan Geothermal Prospect, Alaska.” Unpublished report to City of Akutan, 36p. 10. Stelling and Kent, 2011. “Akutan Geothermal Exploration Project: Geological Analysis of Drill Core from Geothermal Gradient Wells TG-2 and TG-4.” Unpublished report to City of Akutan, 24p. 11. Kolker, Bailey, and Howard, 2011. “The 2010 Akutan Exploratory Drilling Program- Preliminary Findings.” Draft paper submitted to the Geothermal Resources Council for publication/presentation at the GRC Annual Meeting, October 2011. 12. Kolker et al, 2011. “Akutan Geothermal Project: Summary of Findings from the 2010 Drilling Program.” Unpublished report to the City of Akutan, 33p. Geologic Setting Akutan volcano is part of the Aleutian Volcanic Arc, which is Alaska’s most promising setting for geothermal energy. Akutan volcano is one of the most active volcanoes in the Aleutians, with 32 historic eruptions (Simkin and Siebert, 1994). Akutan volcano is a composite stratovolcano with a summit caldera ~1 ¼ mi (2 km) across and 200-1200’ deep (60-365 m; Newhall and Dzurisin, 1988; Miller et al., 1998). Most of the reported eruptions included small-to-moderate explosions from the active intracaldera cone. An initial volcanic hazard review indicated that the proposed geothermal development area was unlikely to be directly impacted by eruption activity consistent with the previous 1500 years, excepting ash fall that might cause temporary closure. The HSBV walls are composed of ~1.4 Ma lava flows, with the SE wall being slightly older and containing numerous dikes. The valley was glacially carved, perhaps during the last major glaciation ending ~8,000 years ago, and potentially reworked during a minor glacial event ending ~3,000 BP (Black, 1975). The HSBV is composed of two linear valleys (the SE-trending Fumarole Valley and the NE-trending valley that contains the hot springs; Fig. 1) joined at right angles, suggesting structural control of glacial flow. Soil geochemical anomalies (Arsenic (As), mercury (Hg), and carbon dioxide (CO2) at the junction of the Fumarole Valley and the HSBV also suggest that the valley junction is structurally controlled and could be an important fluid conduit (Kolker et al, 2010). Hg, As and CO2 occur in anomalously high concentrations near the hot springs as well. Akutan Geothermal Resource Assessment 6 In March 1996, a swarm of volcano-tectonic earthquakes (>3000 felt by local residents, Mmax = 5.1) beneath Akutan Island produced extensive ground cracks but no eruption of Akutan volcano. InSAR images that span the time of the swarm reveal complex island-wide deformation, suggesting inflation of the western part of the island and relative subsidence of the eastern part. The axis of the deformation approximately aligns with new ground cracks on the western part of the island and with Holocene normal faults that were reactivated during the swarm on the eastern part of the island. The deformation is thought to result from the emplacement of a shallow, east-west-trending, north-dipping dike plus inflation of a deep magma body beneath the volcano (Lu et al., 2000). Studies of 3He/4He ratios in Akutan fumarole gasses indicate degassing of relatively fresh near-surface magma (>6 RC/RA; Symonds et al., 2003). This implies that unlike many other composite stratovolcanoes, Akutan’s magmatic plumbing system includes two lateral rift zones, similar to the classic rift zones at Hawaiian volcanoes and elsewhere. These rift zones are aligned along the regional least-compressive-stress axis (John Power, pers. comm.), and serve as active magmatic conduits at shallow crustal depths (Fig. 1). NW- trending rifting appears to be providing the large-scale permeability as well as the magmatic heat source - crucial for the development of an extensive hydrothermal reservoir beneath the volcano. Figure 1. Topographic map of Akutan Island, showing the geothermal project area and pertinent geologic features. Hot Springs Bay Valley (HSBV) is an L-shaped topographic low that lies at the center of the geothermal project area. Akutan village Akutan Geothermal Resource Assessment 7 Geothermics Several thermal springs are located in HSBV, about 6 km from Akutan village (Fig. 1). Five groups of hot springs with about ten vents have been identified, including tidewater springs on Hot Springs Bay beach that are only exposed at low tide. Temperatures range from 129 to 205 °F (54 to 96 °C); and some have been reported as boiling. A fumarole complex (often called the “fumarole field”) exists at the head of HSBV to the west of the hot springs and covers an area of approximately 1600 ft2 (500m2). Motyka and Nye (1988) concluded that the fumaroles are likely fed directly by gases and steam boiling off the deep hot reservoir and that these fluids then mix with cool groundwaters to produce the hot spring waters further down the valley, forming a geothermal system that is at least 2.4 mi (4 km) long. Recent studies of the chemical composition of the fluids confirm that they become extensively mixed with cooler meteoric waters near the surface. Fumarole gas geothermometry indicates that the reservoir fluids attained a temperature of at least 518 °F (270 °C). Cation chemistry from the hot springs and produced fluids indicates that the fluids are re-equilibrating to lower temperatures along the outflow path, with cation geothermometry from the fluids produced by corehole TG-2 providing temperatures of 211-232 °C. The silica geothermometry of 320 °F (~160 °C) indicates that the resource close to the hot springs (probably <1500’ / <500m distance and depth) is likely to be 320-358 °F (160 to 180 °C) (Rohrs, 2011). This is consistent with active silica sinter deposition at the hottest springs. Based on the springs, well TG-2 was expected to encounter a permeable zone with 320-358 °F (160-180 °C) fluid, which it did. The structure(s) controlling upflow of hydrothermal fluids is probably one or more NW-trending normal fault(s). One such mapped fault cuts near-perpendicularly across HSBV (Fig. 2). All of the hot springs are topographically lower than the fault’s surface trace, consistent with geochemical indications that they outflow from an upflow near the fumarole. The fumarole field lies along a parallel linear feature, but no fault has been mapped there. A perpendicular NE-trending fault may control the linear shape of the hot spring locations, but that fault has not been conclusively identified with available data. MT Resistivity The resistivity pattern of the Akutan geothermal prospect has an overall geometry similar to that of most geothermal reservoirs where a low resistivity, low permeability smectite clay caps a higher resistivity, higher temperature, permeable geothermal reservoir. However, the resistivity values of >20 ohm-m within the low resistivity zone at Akutan are much higher than in the smectite zone of most developed geothermal fields. Several models can explain such a pattern, including an unusually high fraction of dense lavas causing weak alteration, or relict alteration that formed at higher temperatures. A localized pattern of alteration near the hot springs is more conventional, with a <600’ (<200 m) thick, 5-15 ohm-m zone that represents a smectite clay cap overlying a higher resistivity geothermal outflow (Figs. 10-13). Therefore, the overall resistivity geometry is consistent with the geochemistry. A tongue of high resistivity at -300 m elevation in Figs. 10-13 trends from the fumarole to the hot springs. This is consistent with a relatively resistive flow path that originates from a >428 °F (>220 °C) upflow near the fumarole to a 358–428 °F (180–220 °C) outflow extending to the hot springs. Unfortunately, steep topography and high winds prevented the MT from accessing much of the prospective area. There are no MT stations over the fumarole area and so a well that targets an interpreted upflow in its vicinity might target only the surface extent of the altered ground and Akutan Geothermal Resource Assessment 8 fumaroles. There are also no MT stations over a large part of the likely outflow path, making it difficult to assess the risk of targeting a well on the accessible part of the likely north flank of the outflow. New Data 2011 1. Temperature Gradient Data 1a. Core Hole Drilling In 2010, two small-diameter temperature gradient (“TG”) core holes were drilled at locations given in Fig. 2. Since the Akutan Geothermal area is roadless, the drilling operations were supported by helicopter. Due to budget constraints, only two of the four planned holes were actually drilled; these are marked with black arrows in Fig. 2. The 2010 exploratory drilling plan was designed to test whether the shallow resource was potentially commercial. Within narrow budget constraints, the wells were designed for long-term monitoring as well as a test of the shallow, accessible targets at the Akutan geothermal field. The hole(s) were completed as temperature gradient wells and available for future monitoring. A detailed report on the drilling operations, P/T survey results, and other data is provided in Kolker et al. (2011). Well “TG-2” was drilled to a TVD (total vertical depth) of 833’ (254 m). It was sited to test the outflow aquifer(s). Between 585 and 587’(178 and 179 m), a highly permeable zone was encountered that flowed geothermal fluid at 182 °C (359 °F). This productive zone was cased and cemented, sealing it off, at which point it cooled to about 329 °F (165 °C). The structure hosting the flowing fluid appeared to be a fractured, highly vesicular, flow margin. Due to the temperature and permeability of the formation at relatively shallow depths, drilling this well was challenging. Although targeted to 1500’ (457 m), the well was terminated due to drilling problems. Well “TG-4” was drilled to the planned TVD of 1500’. It was sited at the southern part of the junction between the two perpendicular valleys, to test the size and extent of the outflow zone. Since well TG-4 did not encounter substantial fluid flow, its location appears to be outside the margins of the outflow zone, vertically or horizontally (or both). However, well TG-4 did encounter an anomalously high shallow temperature gradient, implying close proximity to a geothermal source. Akutan Geothermal Resource Assessment 9 Figure 2. Map of the Akutan Geothermal area, showing the four candidate exploration well locations that were considered for the 2010 program. The two holes drilled in 2010 are marked with black arrows. 1b. End-of-Well Logs After TD was reached, three P/T logs were recorded at 12, 24, and 36 hours after circulation ended for each well. For every run, stops were made at 20 foot stations. Because these surveys were taken so soon after the well was drilled, the temperature readings were still influenced by the cooling effects of fluid circulation. Therefore these represent “unequilibrated” downhole temperatures.In order to predict the equilibrated downhole reservoir temperature, we used the Horner method to extrapolate the measured values to a longer period. The end-of-well elevation vs. temperature plot generated from extrapolated Horner values is shown in Fig. 3. The MRT reading from the flowing zone is also shown as a purple dot. Akutan Geothermal Resource Assessment 10 Figure 3. Estimated equilibrated Temp v. depth plot for TG-2, based on Horner extrapolations of downhole survey data (see Fig. 11 and text for details). Both wells show very high shallow temperature gradients, which is consistent with their proximity to the shallow outflow zone. Following production, TG-2 shows a drop in temperature occurring just above the casing shoe at 603’(181 m), corresponding to the hot fracture zone between 585 and 587' (178 and 179 m) that was cemented in. The apparent cooling is likely the result of drilling fluid and cement injected across that entire area. TG-4 shows a relatively rapidly increasing temperature gradient until ~900’(274 m),transitioning to a slowly increasing temperature gradient from 900’-1500’. The fact that there was no temperature reversal and that the gradient continues to increase suggest there could be a deeper, hotter aquifer below 1500’(457 m) that was not penetrated by drilling. An injection test performed on well TG-4 suggested that the well has generally poor permeability. 1c. Equilibrated TG Logs While the end-of-well surveys were conducted with a memory tool, it was not possible to use a memory type tool for the post-completion logging due to the small inner diameter of the Akutan TG wells (inner diameter > 1.5 inches / 3.81 cm). Due to this and other unique conditions of Akutan TG wells (high -1600 -1400 -1200 -1000 -800 -600 -400 -200 0 200 0 100 200 300 400 500 Elevation (feet ASL)Akutan TG Well Temperatures (F) TG2 TG4 Boiling Curve 0.1% NCG Flowing MRT TG2 Akutan Geothermal Resource Assessment 11 temperatures at shallow depths, remoteness of the wellsites, among others), thermistor-type temperature logging equipment was used, and downhole pressures were not recorded. The survey for TG-2 was completed on May 22, 2011. Results from the equilibrated survey are shown with the three build-up surveys in Fig. 4. Figure 4. Equilibrated temperature profile for TG-2, plotted with the three end-of-well heat-up surveys. The end-of- well surveys were taken 12 hours. 24 hours, and 36 hours after circulation; the equilibrated profile was obtained 9 months later in May 2011. The new temperature profile shows a distinctly different shape from the end-of-well temperature build- up profiles. Among the new features to note are: (1) The well was bleeding while the log was run, resulting in a minor steam or two-phase section in the upper 60-70’ (20-25 m). (2) Apparent cooling of the well since shut-in is noticeable in the upper 400’ (122m). This probably reflects a trickle of water downflowing from around 200’ (61m) MD and exiting into the formation at about 415’ (126m) MD. It can only be a trickle of water because the water is heating up as it flows down behind the casing. (3) The highest temperatures occur in the permeable zone near 585’ MD (415’ / 126m elevation), with a temperature reversal of about 9 °F (5 °C) below the permeable zone to the bottom of the well. 0 100 200 300 400 500 600 700 800 900 0 50 100 150 200 250 300 350 400Elevation (feet ASL)Akutan TG2 Temperatures, oF 12h 08-2010 24h 08-2011 36h 08-2011 May-11 Akutan Geothermal Resource Assessment 12 The new data also shows that the permeable zone at 585’ MD (415’/ 126 m elevation) has fully recovered in temperature. Notably, the static temperatures measured in this permeable zone are about 338 °F (170 °C), which is lower than the 359 °F (182 °C) temperature measured in this zone when the well was flowing. Since the MRT reading does appear to be correct based on silica geothermometry, this implies that the well was drawing in higher temperature fluids when it was producing. A temperature survey was run in TG-4 by the City of Akutan crew on May 10, 2011 (Fig. 5). Figure 5. Equilibrated temperature profile for TG-4, plotted with the three end-of-well heat-up surveys. The temperature profile from the equilibrated survey differs very slightly from the end-of-well temperature profile. The new profile shows that the top 800’ (244 m) of well TG-4 heated up slightly, but the bottom temperatures remained extremely close to those measured during the end-of-well surveys. This is not surprising in light of the fact that that well was relatively impermeable and exhibits a temperature profile that shows heating primarily from conduction for the upper 800’ (244 m). By contrast, the bottom of the hole is approaching an isothermal gradient. This suggests that the conductive heating is from the side (i.e., from a shallow outflow zone at some lateral distance), not from a hot aquifer below. -200 0 200 400 600 800 1000 1200 1400 1600 0 50 100 150 200 250 300 350Elevation (feet ASL)Akutan TG4 Temperatures, oF 12h 08-2011 24h 08-2011 36h 08-2011 May-11 Akutan Geothermal Resource Assessment 13 1d. P/T Data Analysis Although TG-2 flowed geothermal fluid at 359 °F (182 °C) during drilling, the equilibrated temperature logs show a maximum temperature of 338 °F (165 °C) with a reversal at the bottom of the hole. This implies that the 359 F fluid was not circulating in the immediate vicinity of TG-2 but rather was “pulled in” from elsewhere due to the pressure drop caused by flowing the well. A likely scenario is that the productive subhorizontal fracture at 585’ (178m) in TG-2 is connected to a subvertical fracture dipping west (see Fig 11). When the subhorizontal fracture was produced, the subvertical one became a temporary conduit for fluids in the outflow zone. It is unlikely that the source of the 359 °F (182 °C) fluid is directly below Well TG-2 because of the temperature reversal recorded in the most recent log. A comparison of the static temperature profiles in TG-2 and TG-4 shows the difference between the shape of a convectively heated outflow profile in TG-2, and a conductively heated temperature profile in TG-4 (Figure 6). Also, the temperatures in the upper 800’ (254 m) of TG-4 are generally lower than in TG- 2, indicating that TG-4 is further from the shallow outflow path. No strong conclusions can be drawn from the temperature profiles as to whether additional high temperature permeable zones underlie either well, but it appears unlikely based on the shape of the bottom of both well profiles. Figure 6. Equilibrated well profiles for both Akutan TG wells, shown with a boiling point with depth curve for water with 0.1% non-condensable gas content. -200 0 200 400 600 800 1000 1200 1400 1600 150 200 250 300 350 400 450Elevation (feet ASL)Akutan TG Well Temperatures May 2011 (oF) TG2-05-2011 TG4-05-2011 Boiling 0.1% NCG Akutan Geothermal Resource Assessment 14 The 359 °F (182 °C) temperature measurement during drilling of TG-2 does appear to be correct based on silica geothermometry. If the well was drawing in higher temperature fluids when it was producing, this suggests that TG-2 was drilled on the margins of a more permeable and hotter outflow path. The higher temperature fluids drawn into TG-2 during the flow test suggest that the production zone is in proximity to the higher temperature zone but that it has a relatively low permeability connection to this zone. The slight temperature reversal of about 9 °F (5 °C) below the permeable zone is consistent with the geologic model that the thermal features in HSBV represent a confined lateral outflow from a geothermal reservoir located further west, or possibly north. The temperature gradients for Akutan wells TG2 and TG4 vary widely, but compared to the continental average of 1.65 °F/ft (30 °C / km; Fig. 7) they are very high above 600’ (244 m). This suggests that both are within proximity of a very shallow outflowing resource. Figure 7. Temperature gradients, in degrees Fahrenheit per foot, for both Akutan TG wells. The average continental geothermal gradient of 1.65 °F/ ft is shown for comparison. The outlier data point at ~410’ depth can be ignored as it reflects a transition between the part of the well affected by a small amount of downflowing water and the part unaffected, and thus does not represent an accurate temperature gradient. 0 200 400 600 800 1000 1200 1400 1600 024681012141618Elevation (feet ASL)Akutan Well Temperature Gradients (T2-T1, oF) TG2 TG4 Average continental Akutan Geothermal Resource Assessment 15 2. New fluid chemistry and geothermometry 2a. Sample Collection and Data Sources Two fluid samples were obtained from well TG-2, with the first obtained from the entry zone at 585’- 587’ (178-189m) measured depth (MD) during a well discharge. This production zone was subsequently cased off. A second flow test of the well obtained samples from production zones between 603’ (184m) and 833’ (245m) MD, which was the completion depth of the core hole. Because TG-4 encountered poor permeability conditions, a sample of the fluids in the wellbore was obtained by flowing with an air assist. The MD of TG-4 was 1500’ (457m) and had a cemented casing at 596’ (182m) MD. Therefore the data obtained during the discharge of the wells vary in quality. New gas chemical data are from samples obtained from the fumaroles in 2010. All other analyses used in geothermometry calculations and chemical modeling were obtained from past reports (fluid analyses from the hot springs, non-condensable gas analyses from the summit fumarole and from the hot springs – see p. 4 for sources). 2b. Chemistry Geochemical data were interpreted using a combination of binary and ternary diagrams and geothermometry gas plots. The data set used to interpret the reservoir conditions consists of all of the data obtained at Akutan. Data and interpretations are provided in full in Rohrs (2011). The chemical analyses of the hot springs water shows that they are derived from a dilute, near-neutral Na-Cl reservoir brine. The Akutan hot springs show slightly elevated HCO3- and SO4 concentrations, suggesting mixing along the outflow path with dilute, steam-heated near-surface waters. Hydrogen and oxygen isotopic data shows that the hot spring waters are derived from local meteoric water. The chemistry of the fumarole gasses demonstrates a strong magmatic affiliation. There is no evidence from the gasses that the reservoir water has mixed with air-saturated fluids along an outflow path. Compared to many geothermal systems, Akutan displays enriched N2 concentrations, which in some cases would raise concerns with regards to acid or vapor-cored conditions in the reservoir. However, the other gas plots show that the gases are well-equilibrated and likely to be derived from a high temperature neutral chloride reservoir. In addition, the gas concentrations in the flank fumaroles imply that some fraction of gas is derived from equilibrated steam, indicating the presence of a localized steam cap in the reservoir. The chemistry of the fumaroles are consistent with an equilibrated geothermal system associated with an andesitic stratovolcano (Giggenbach, 1991). In comparison, the gas from the summit fumarole originates from a more oxidizing environment and exhibits high H2S concentrations. These all suggest a magmatic affiliation for the summit fumarole steam. 2c. Geothermometry The hot spring and well discharge samples are well suited to chemical geothermometry using the silica and Na, K, Ca, and Mg concentrations of the fluids. The estimated temperature of last equilibration along the outflow path suggests that the fluids have equilibrated at ~338 °F (~170 °C) and ~392 °F (200 Akutan Geothermal Resource Assessment 16 °C) for the two samples from TG-2. This temperature is similar to the estimated entry temperature of 359 °F (182 °C) at 585-587’ (178-179m) MD in well TG-2 (Kolker et al, 2010). Cation concentrations in hot spring and well discharge analyses show that the springs and well fluids are mixed or partially equilibrated fluids. This is commonly observed along outflow paths where the fluids are re-equilibrating to lower temperatures and mixing with near surface waters with elevated Mg concentrations (Rohrs, 2011). The data from hot spring HS-A3 and the entry at 585’ (178 m) MD in core hole TG-2 suggest that the fluids originate in a deeper reservoir with temperatures in the range of 428-464 °F (220-240 °C). This compares to a temperature of 412 °F (211 °C) estimated from the Na-K-Ca geothermometer for the well discharge. Geothermometers that apply Na, K, Ca, and Mg concentrations tend to partially re-equilibrate to lower temperatures in the outflow zone, and so the deep reservoir temperature is likely to exceed 464 °F (240 °C; Rohrs, 2011). Geothermometry estimates from flank fumarole gasses exhibits very good consistency, indicating an origin from a mature, equilibrated neutral chloride reservoir. The gas geothermometry consistently suggests reservoir temperatures of 518-572 °F (270-300 °C; Rohrs, 2011). 2d. Geochemical Model The new geochemical data set confirms the previous interpretations of the resource distribution in HSBV. The hot springs represent a shallow outflow from a high temperature neutral chloride reservoir that exists further west. The chemistry of the hot springs indicates that they have experienced significant mixing with cooler, dilute near surface meteoric waters. Because the fumarole gases show little evidence of mixing with air-saturated waters, the upflow zone is likely to lie near the fumaroles. Also, gas grid plots indicate that the fumaroles contain a component of equilibrated steam, suggesting the possibility that a localized steam cap overlies the deeper geothermal reservoir. Geothermometry of the well discharges and the fumarole gases indicate a likely deep reservoir temperature of at least 464 °F (240 °C) based on Na/K geothermometry, with temperatures possibly as high as 572 °F (300 °C) in the upflow based on gas geothermometry. The geochemical data do not provide any constraints on the reservoir boundaries to the west nor on the reservoir volume within the outflow area (Rohrs, 2011). The non-condensable gas data from the fumaroles suggest that a steam cap may overlie the deep brine reservoir. The chloride hot springs in HSBV represent shallow outflow from the reservoir. The outflow becomes diluted by mixing with cool meteoric waters, especially in the near surface environment (Rohrs, 2011). Thus, the geochemical data are very consistent with the geochemical outflow models suggested by Kolker et al. (2009). 3. Core data 3a. Overview Composite logs from Akutan TG wells TG-2 and TG-4 of the bulk lithologies, alteration mineralogy, and temperature data are provided in Appendix A. Full lithologic logs were recorded at the wellsite during drilling and are provided as an Appendix in the Summary of Drilling Findings (Kolker et al., 2011). The core was then sent to Western Washington University (Bellingham, WA) for detailed laboratory analysis. Akutan Geothermal Resource Assessment 17 The goal of the laboratory analysis was to determine the hydrothermal history of the HSBV. Core samples were selected based on zones of interest from drilling records, core photographs, and complete coverage of the depth of core. Determination of specific mineral species was conducted through X-ray Diffraction (XRD) analysis, Scanning Electron Microscopy (SEM), and petrographic observations. Quantitative permeability studies of the core were not conducted, however qualitative observations about the permeability of the field by visual observations of the core were recorded. Finally, compositional analysis of 19 bulk rock samples were conducted by X-ray Fluorescence (XRF) at the Geoanalytical lab at Washington State University in Pullman, WA. Methodologies for above studies, detailed results, and discussions are provided in Stelling and Kent (2011). 3b. Rock Types and Primary Mineralogy There are four main lithologies present in the Akutan core: basalt, andesite, ash tuff, and “lithic basalt.” The most common lithology in the core is basalt lava. These flows appear to be subareal in nature and contain plagioclase, clinopyroxene, rare olivine and primary apatite. The ash tuffs are fine grained rocks lacking phenocrysts. Groundmass phases are plagioclase microlites, glass, and alteration minerals (see below). In TG-2, these units are <3’ (1 m) thick. In TG-4, which is ~2 miles (3.2 km) closer to Akutan Volcano, similar units are as thick as 60’ (18 m). The rocks provisionally named “lithic basalt” were a puzzle during the on-site evaluation, and remain enigmatic. The lithic basalt is composed of multiple different rock types, suggestive of some sort of debris flow deposit, yet the matrix between the clasts is crystalline, indicating a magmatic origin. At this time, the origin of this lithology is unknown. 3c. Secondary Mineralogy, Mineral Paragenesis, and Hydrothermal History A graphical representation of secondary mineralization and clay replacement is presented in Stelling and Kent (2011). The rocks in general appear to be only weakly altered. As a result of the increased porosity near lava flow tops, these regions tend to be more altered and more readily brecciated than the main body of the lava. Heavy Fe-oxidation was observed between flow layers. Alteration minerals occurred interstitially, in fractures, in vesicles, and in contact zones. Alteration assemblages in both wells are dominated by chlorite, zeolites, epidote, prehnite and calcite, and this alteration appears to have happened multiple times in both wells. The presence of adularia in specific locations in both wells indicates higher temperature and permeability conditions existed at some point in the past. The presence of kaolinite in TG-2 indicates argillic alteration with lesser extent and intensity. Illite was identified in both wells, although much more sparsely in TG-2. Within the most recent propylitic alteration event in TG-2, the sequence of zeolite formation shows a classic trend toward higher temperatures with depth. It is likely that this trend will continue below the base of the well (833’ / 254 m). Figure 8 shows that some higher-temperature minerals (illite, epidote, prehnite, wairakite and adularia) occur in regions that are currently much colder than expected for these minerals. This suggests that the TG-2 region underwent higher temperature alteration in the past. The presence of these higher-temperature minerals at unexpectedly shallow depths further suggests that a significant portion of this older alteration sequence has been removed through erosion, possibly Akutan Geothermal Resource Assessment 18 glacial. Overprinting of these minerals by lower-temperature alteration assemblages indicates the sampled region has since returned to a lower-temperature alteration regime with reduced permeability. Figure 8. First occurrence of indicator minerals with depth in core from Akutan well TG-2. Horizontal arrows indicate formation temperature ranges for each mineral. Dashed lines indicate published values; solid lines indicate the most commonly reported minimum temperatures. The pattern of alteration in TG-4 is more complex than TG-2 (Fig. 9). That the depositional history of TG- 4 includes multiple alteration events does not necessarily mean that the two wells have had significantly different thermal histories; rather, the most recent alteration event may have been stronger in the TG-2 region, overprinting more completely the alteration sequence observed in TG-4. Akutan Geothermal Resource Assessment 19 Figure 9. First occurrence of indicator minerals with depth in core from Akutan well TG-4. Horizontal arrows indicate formation temperature ranges for each mineral. Dashed lines indicate published values; solid lines indicate the most commonly reported minimum temperatures. Akutan Geothermal Resource Assessment 20 Comparing the observations made in the two drill cores provides a basic sequence of alteration for the HSBV geothermal outflow zone. Both cores show an alteration sequence progressing from an early propylitic event, a narrow band of adularia-bearing propylitic alteration, followed by a later propylitic event. The trend from moderate propylitic to high-temperature adularia-forming alteration and back to moderate propylitic indicates that the shallow portion of the HSB field has reached its thermal peak and has cooled moderately. Additionally, many of the higher temperature minerals occur at depths much shallower than reported in other geothermal fields. Thus is it likely that 1) this region was hotter than it is currently, and 2) the uppermost portion of the rock column has been removed and these rocks have risen to their modern shallow depths. 3d. Permeability and Porosity of Well Rocks The primary lithologies do not lend themselves to high primary permeability. The abundance of isolated vugs filled with secondary minerals indicates that fluid flow through microscopic intergranular networks has been important, but flow rates are likely very low. Vug filling is especially common in fine-grained, detrital deposits (e.g., ash tuff), but clay alteration and fracture mineralization by carbonates and zeolites reduces permeability in these rocks. The primary fluid pathways appear to be associated with brittle fracturing and lithologic contacts, based on the abundance and degree of alteration and secondary mineralization. Very fine grained deposits (ash tuff) lack large crystals that would add structural control over the fracture patterns. As a result, these rocks are prone to planar fractures at prescribed orientations (30o, 45o, 60o and 90o). The majority of these fractures have some sort of secondary mineralization associated with them, and some of the larger fractures are deeply altered to a variety of clays and other minerals. Because the tuff is more susceptible to clay alteration, these fractures can seal before major secondary mineralization becomes intense. However, these units are not very thick in the wells, so may not have extensive control over the overall fluid flow in the reservoir. Permeability may locally increase at the top of lava flows where vugs in vesicle-rich flow tops may collapse, but this was not observed in the core. Some heterogeneous lithologies (e.g., “lithic basalt,” see below) contain entrained clasts of older material . Fluid flow within these lithologies are concentrated and directed around the entrained clasts, which would likely result in moderately increased permeability compared to intergranular flow. The occurrence of the mineral adularia helps to elucidate the permeability. Although adularia occurs in all lithologies in the HSB cores, the restriction of adularia to fractures highlights the importance of secondary permeability, as it does in many fields worldwide. Adularia is strongly associated with zones that once had high permeability but each occurrence of adularia in the core is in veins that are now thoroughly sealed by mineralization. Therefore, the waxing of a higher temperature system and subsequent waning has apparently reduced the permeability in the HSBV outflow system. No evidence for large scale structures were encountered in Akutan geothermal wells. A number of brecciated zones were observed in TG-4, but most were “sealed” with secondary mineral deposits and therefore probably do not represent active faults. Minor slickensides observed in cores could be related to a possible normal fault on the SW side of the valley near TG-4. Akutan Geothermal Resource Assessment 21 Resource Conceptual Models Several conceptual models of the Akutan Geothermal Resource were presented in an earlier report (Kolker et al, 2009). The acquisition of new data in 2010-2011 are consistent with the same basic upflow – outflow model. The most important change in the conceptual model assessment is the increase in confidence in the fumarole as the locus of a benign reservoir upflow that would be a suitable target, based on the promising new gas geochemistry analyses. The new data significantly reduces the probability that an economic reservoir would be found in HSBV. The location of the high permeability outflow path is still only loosely characterized by two models, although the upflow seems more closely connected to TG-2 than to TG-4. The alternative outflow pathways continue to be either along the HSBV or along a northern trajectory from the fumaroles to the hot springs. These two alternatives are explored in Figs. 10-13. Both conceptual models are based on the notion that the Akutan geothermal system is a single resource comprised of two distinct features: a high-temperature (>500 °F / >240 °C) upflow zone, and a lower-temperature outflow aquifer (~360-390 °F / 180-200 °C), as suggested by chemical data. Figure 10. Map view of “Conceptual model 1,” showing outflow along HSBV. Isotherm contour placement is based on downhole temperature data, chemical geothermometry, hot springs and fumarole locations, and MT resistivity data (shown here at -300 m (~984 ft) depth). Profile line “CM1” corresponds to Fig. 11. Akutan Geothermal Resource Assessment 22 Figure 11. Profile “CM1,” as shown in Fig. 10. This model shows outflow along HSBV. Isotherm contours based on downhole temperature data, chemical geothermometry, hot springs and fumarole locations, and MT resistivity data (shown here as 3D inversion model). Conceptual model ‘CM1’(Figs. 10 and 11) shows a high temperature resource upflowing beneath the fumarole field, and cooling along an outflow path that follows the L-shaped path of HSBV. The upflow must be some lateral and vertical distance from well TG-4, since no trace of conductive heating from a deep source was observed in the temperature profile of TG-4. Also, for this model to fit the observed downhole temperature profiles in both wells, the outflow along HSBV can only be very thin (vertically constrained low-permeability) and restricted to the shallow subsurface. Because the 359 °F (182 °C) flow during the flow test completed while drilling is higher than the measured static temperature, because there does not appear to be a downflow that would reduce the temperature of this zone from 359 °F (182 °C) to 338 °F (165 °C) when the well is static, and because the temperature reversal in TG-2 below this zone makes upflow from below the well unlikely, the produced higher temperature fluid appears to have been “pulled in” laterally from a nearby source. This could be related to the westward-dipping fault near TG-2 shown in Figs. 11 and 13. The rapidity with which this hotter fluid was drawn in during such a short test implies that the 338 °F (165 °C) permeable zone in TG- 2 must be restricted in volume and at a higher natural state pressure than the 359 °F (182 °C) adjacent reservoir. Akutan Geothermal Resource Assessment 23 Figure 12. Map view of “Conceptual model 2,” showing outflow beneath the mountain to the north of HSBV. Isotherm contour placement is based on downhole temperature data, chemical geothermometry, hot springs and fumarole locations, and MT resistivity data (shown here at -300 m (~984 ft) depth). Profile line “CM2” corresponds to Fig. 13. Conceptual model 1 ‘CM1’ does not resolve the location of a hotter outflow resource of 360 -392 °F (180- 200 °C), for which there is a substantial amount of geochemical evidence. Therefore, an alternative model is proposed called ‘CM2’ (Figs. 12 and 13). In CM2, the shallow outflow path takes a northerly trajectory from the fumarole to the ENE towards the hot springs, circumventing HSBV altogether. This model appears more likely based on several lines of reasoning: 1) the temperature profile for TG-4 shows no evidence for being along an outflow path, implying that outflow feeding the hot springs is laterally distal; 2) a low-resistivity clay cap appears to form a dome pattern around the northerly outflow path, which is consistent with the interpretation that the HSBV is near, but not in, the main outflow path of geothermal fluids (Figs. 10 and 12); and 3) the isotherm contours on the CM2 profile (Fig. 13) are slightly more typical of an outflowing geothermal system. Akutan Geothermal Resource Assessment 24 Figure 13. Cross section of profile line “CM2,” as shown in Fig. 12. This model shows outflow beneath the mountain north of HSBV. Isotherm contours based on downhole temperature data and chemical geothermometry, MT resistivity data, hot springs and fumarole locations, and fault lines. Both models suggest that producing the outflow resource would be very risky, both because of the generally low permeability expected based on several lines of reasoning and also because there is no well-developed clay cap to indicate that a large-very permeable reservoir volume at ~360-390 °F (180- 220 °C) exists under HSBV. The lack of widespread surface alteration, geochemical, and ground temperature anomalies (Kolker et al., 2009) in HSBV are consistent with this interpretation. Additionally, the chemical composition of the hot springs fluids suggests that outflow fluids become extensively mixed with cooler meteoric waters near the surface, raising concerns about cold water influx into the outflow system with production. While the conceptual models of the outflow resource have downgraded its potential for development, geochemical data from the fumaroles significantly upgrades the fumarolic area as a drilling target. The fumarole data suggest that the flank fumarole field lies in fairly close proximity to an upflow zone from the reservoir, that a steam cap may overlie the upflow, and that reservoir temperatures could approach 570 °F (300 °C) within the upflow. The deep reservoir probably consists of a brine liquid capped by a small two-phase region (steam cap) (Rohrs, 2011). Resistivity data suggest that the upflow reservoir is situated in brittle rocks, implying propylitic alteration regime and a good possibility of high permeability. Akutan Geothermal Resource Assessment 25 Future Drilling Targets The two types of resource targets represent two development options for the AGP: 1) The upflow resource has a high exploration risk, but potentially low development risk and high output capacity. 2) The outflow resource has potentially easier access but higher exploration risk, and presents several development risks, including a lower output capacity per well. The highest priority target for future drilling is at the upflow resource, due to the following factors: (1) Evidence of commercial-grade permeability due to the presence of fumaroles that are chemically connected to a neutral-chloride reservoir with a steam cap. (2) The fumarole gas geothermometry indicates that the source fluids are likely to be being equilibrated to a temperature of >572 °F (>270 °C). These temperatures could exist directly beneath the fumaroles. Alternatively, if the upflow originates further west, the fumaroles may mark the location where the outflow first encounters boiling conditions. In this case, temperatures beneath the fumaroles could be in the range of 428-464 °C (220—240 °C). (3) As the highest-enthalpy target, the fumarole area could be expanded to meet additional power demand if it should ever present itself (e.g., new industrial capacity, larger scale secondary use, etc.) The exploration data suggests that the likely upflow location is in the general vicinity of the fumarole. However, the fumarole is located ~1150’ (~350 m) up a very steep hillside, posing access limitations. Hence, an important issue is the trade-off between the cost and practicality of constructing a pad closer to the fumarole and drilling further directionally. A small rig may only be able to achieve a very modest directional offset, but a rig capable of greater directional offset would be several times more expensive. Two alternative surface locations have been sited to target the high priority upflow zone. Regardless of the pad location, the well should be targeted to at least 4500 ’ (1350m) MD and preferably to 6000’ (1800 m) MD. The first, preferred alternative “A” is located near the fumarole field (Fig. 14). This is closest to the high temperature upflow zone. Access to this location could be via a road running east-west which skirts the mountain to the north of HSBV. Such a road appears to be buildable at less than 3 miles (6 km) from Hot Springs Bay access, but would require dock facilities to be built at the beach. However, this access option raises the question of transmission to Trident and Akutan Village. It also raises questions about volcanic hazards (see ‘Risks’ section, below). The second, less preferable alternative is located in the Fumarole Valley >2/3 mi (~1 km) southeast of the fumarole field (Fig. 14). A directional well drilled from this location (called “Well -1” in past reports) beneath the fumarolic features or to the north beneath the local resistivity high may intersect the upflow zone. From pad location “B”, the margins of the upflow resource would be targeted via directional drilling and the outflow resource could be targeted via vertical drilling. However, limitations Akutan Geothermal Resource Assessment 26 on directional drilling may not allow the target to be reached from pad location “B.” Hence, this wellpad location is riskier than alternative “A.” A determination on this issue should be solicited from a qualified geothermal drilling engineer before a final decision is made. Figure 14. Map showing possible wellpad locations to target the Akutan upflow resource (blue triangles). Each pad could host two wells – a directional well aimed towards the fumarole field (knobbed black line), and a vertical well. Possible road alignments to the wellpads are shown in red. Also shown are the three sections recently selected by the Akutan Corporation for subsurface ownership rights in a land swap agreement with the Aleut Corporation. Should drilling at sites A and B fail, or if developing the upflow resource is not possible, subsequent pad locations could be sited to target other parts of the outflow zone. A 380 to 428 °F (180 to 220 °C) outflow resource target about 2200’ (800 m) to the northwest of TG-2 might be preferred if its higher targeting risk and lower generation per well due to its lower temperature were sufficiently offset by lower drilling and access cost. Prior to generating outflow targets, additional exploration activities should be undertaken to target the hottest and most permeable part of the outflow resource. This would likely require additional subsurface imaging work using one or several geophysical techniques. It would also probably require additional slimhole drilling and/or deepening of TG-2. Since the casing in TG-2 was not cemented in place, it could be retrieved it and the well deepened by 1000’ (300m) or more. Akutan Corporation Subsurface Selections 1 36 12 A B Akutan Geothermal Resource Assessment 27 Capacity Assessment As summarized by Glassley (2008), different approaches and methodologies to geothermal resource capacity assessment have given rise to a broad range of results that are not directly comparable. Hence, the outcomes of resource assessments are sensitive to the methodology and assumptions employed in the analysis, and different studies often produce widely different estimates of a resource. The assessment of reserves by analogy used in previous Akutan reports is updated here based on the results of the wells and the fumarole gas geochemistry. The approach applied here is to use analogies to a few published examples in order to highlight important similarities and differences with respect to Akutan. Resource Existence and Size Resource risk assessment approaches commonly divide the assessment into two parts; 1) an assessment of confidence in the existence of a resource as a percent probability, and, 2) assuming the resource exists, an assessment of its size, usually as a statistical distribution (e.g. Newendorp and Schuyler, 2000). The probability of existence is sometimes restated as the probability of exploration success; i.e., the probability that an exploration drilling program would discover at least one economically productive well. In many published geothermal resource assessments, the assessment of existence is often not explicitly evaluated but nominally included in the size distribution, for example, in the Western Governors’ Association (2006), Clean and Diversified Energy Initiative Geothermal Task Force Report. In this report, many geothermal prospects in the western USA with poorer indications of temperature over 220 °C and much lower surface heat flow than Akutan are assessed as having over 20 MW potential. Confidence in Resource Existence The most common method of estimating the probability of existence for a resource is to have a group of experts review the available data and, based on analogous experience with other geothermal prospect areas, estimate the confidence (as a probability) that the necessary components of a resource exist together. For volcanic prospects that have hot springs with cation geothermometry similar to Akutan’s and a non-magmatic fumarole, few failure cases exist in which the most attractive target was drilled. With respect to the earlier assessments of Akutan, the very minor magmatic indications and excellent gas geothermometer estimates of resource temperature have increased confidence in the existence of a high enthalpy resource. The numerous geothermal success cases differ in detail, particularly with respect to the geology and very dilute chemistry characteristic of Akutan. For example, in the Americas there are several developed geothermal fields in volcanic systems with different geologic settings but broadly similar liquid and/or gas geochemistry. The 572 °F (270 °C) San Jacinto Field in Nicaragua has 10 MW installed and 72 MW under development. It has roughly analogous fumarole gas geochemistry and a similar area of intense alteration, although the resistivity of its clay cap is much lower, more like a conventional geothermal field. The 320 to 350 °F (160 to 175 °C), ~40 MW Casa Diablo field at Long Valley (Sorey et al., 1991) and the 320 to 360 °F (160 to 180 °C), 45 MW Steamboat Springs Field near Reno (Mariner and Janik, 1995) have liquid chemistry similar to Akutan, but again, a lower resistivity clay cap. At Akutan, the combination of a non-magmatic flank fumarole with excellent gas geothermometry over 518 °F (275 °C), a trend in cation geothermometry to >428 °F (>220 °C), silica geothermometry over 320 °F (160 °C) with Akutan Geothermal Resource Assessment 28 sinter deposition proven to exist in the subsurface by a well support the existence of a convecting geothermal resource on Akutan with a high confidence of 80%. Probable Resource Capacity The capacity of the geothermal resource at Akutan in terms of electrical power can be assessed using analogies, both the rough comparisons to the prospect estimates provided in the Western Governors’ Association report and the analogs to the 20 to 72 MW San Jacinto development and the 40 to 45 MW Casa Diablo and Steamboat Springs developments. Because of the dilute outflow chemistry and low permeability relict alteration at Akutan, handicapping the Akutan likely 320 to 358 °F (160 to 180 °C) outflow resource by 75% relative to these developed reservoirs would be reasonable, giving an analogous low temperature resource capacity estimate of 15 MW with a 66% probability. Because a high temperature resource very likely exists, a more optimistic capacity estimate for the entire system would be like San Jacinto, 10 to 72 MW with a 66% handicap because of its high clay cap resistivity and difficult access, this results in a risk weighted estimate of about 20 MW. Using the Western Governors’ Association report assessments as analogs, an assessment as high as 100 MW seems reasonable. An alternative approach An output capacity for the 359 °F (182 °C) fluids produced by TG-2 was estimated in 2010 based on the flowing temperature of the well and assumptions about flow rate (Kolker et al., 2011). Based on then- available information, it was estimated that a production well drilled at or near the TG-2 site could produce 1.34 MW up to a maximum of 2.38 MW. However, recent data and analyses including the stabilized temperature curves, alteration mineralogy from cored rocks, have supported revisions of the earlier assumptions used for estimating wellhead flow capacity were optimistic. Monte Carlo Heat-in-Place Option In previous reports on Akutan (Kolker, 2010), the heat-in-place method has been outlined but it has not been formally applied. Initially developed by the USGS for rough regional estimates (Muffler, 1979), more elaborate Monte Carlo versions of the method have recently been adopted by stock exchange regulators in Australia and Canada as a standard for publishing geothermal reserves (Lawless, J., 2010). Despite its common use by geothermal investors, as detailed by Garg and Combs (2010) and more generally considered in the context of other methods by Grant and Bixley (2011), Monte Carlo heat-in- place approaches are commonly misleading and difficult to validate. If such an analysis is needed to meet a reviewer’s request, the City of Akutan could consider employing a large consultancy that routinely provides such analyses to meet regulatory needs, like GeothermEx or SKM. Akutan Geothermal Resource Assessment 29 Resource Risks The major volcanic hazard posed to a geothermal development on Akutan is ash fall. The modern volcanic complex forms the western half of the island, and future eruptions are unlikely to affect the eastern portions of the island (Ancestral Akutan), including HSBV. Destabilization of the fumarole area, at an elevation of ~1150’ (350 m; see Fig. 1), may generate debris flows, and such deposits are seen in the valley floor. According to the hazards map of Akutan, it is possible that the entire HSBV could be inundated by cohesive lahars associated with small-scale slope failure(s), but not likely. Another possible but unlikely hazard is a pyroclastic flow near the fumarole field (Waythomas et. al., 1998). Re-injection beneath the surface is the most environmentally responsible means of disposing of the produced fluids. Re-injection also supports reservoir pressure. Normally the fluids would be injected back into the reservoir because this is where adequate permeability exists. Because of the possibility that HSBV is fault-controlled, reinjecting the fluids into the shallow aquifer incurs a high risk of pre- mature thermal breakthrough to the producing wells. This risk can be investigated through pressure- interference and possibly tracer testing as additional delineation wells are drilled. Calcite scaling in the production wells and silica scaling of the production pipeline system and injection wells are both possibilities for the Akutan system. However, gas levels are likely to be moderate and calcium and silica concentrations are low, suggesting that scaling risks should be low. Silica is unlikely to achieve a significant level of supersaturation should the fluids be produced to a binary power plant. Also, the risks of silica precipitation can be mitigated through pH modification of the produced waters. In addition, the potential for producing acid corrosive fluids at Akutan is very low. The hot springs and discharge waters demonstrate that the reservoir hosts a near neutral chloride reservoir. Further assessment of the risks would require the acquisition of additional well performance data, such as interference testing, and additional geochemical samples from the production and injection zones. Upflow Development Risks Two key questions remain unresolved concerning the deep high temperature (“upflow”) resource. The first is the location of the upflow to the system. The second is the volume of the high temperature resource. Resolution of these questions would require additional drilling and possibly the acquisition of additional resistivity data near and to the west of the fumaroles. A reasonable minimum size at a 10% confidence level would be the area covered by the fumaroles and gassy alteration. If this is taken to be roughly 1500 ft2 (0.5 km2) then using the base case numbers for power density of 15 MW/km 2, a reasonable minimum for expected capacity of the upflow is 8 MW after Grant and Bixley (2011). The risk of volcanic hazards should be carefully investigated if the wellpad were to be sited near the fumarole field on the flank of Akutan volcano, as that location lies on a possible path of pyroclastic flows from Akutan volcano (Waythomas et al, 1998). Fumarole gas contents are 3.5 to 4 wt. %, which do not present an obstacle to development. A well drilled beneath the fumarolic complex is also at a low risk of encountering acid fluids, because the gas chemistry is indicative of a neutral chloride upflow to the geothermal system. Akutan Geothermal Resource Assessment 30 Outflow Development Risks Other than the general risks mentioned above (injection breakthrough, scaling/corrosion, etc), the most important development risk of the outflow resource discovered by core hole TG-2 is permeability and/or volume limitations. Neither of the coreholes encountered clay alteration characteristic of a well- developed smectite-rich argillic caprock. Perhaps the argillic alteration did not develop because of dense lavas, or higher rank alteration that does not retrograde, or it never became well-developed, or perhaps it was eroded off. Resistivity profiles of the HSBV also suggest only a very thin conventional clay has formed over the outflow system, close to the hot springs (Kolker et al., 2009). The rocks in general appear to be only weakly altered. This implies that HSBV does not host a substantial volume of hot fluid, making it a risky development target. Mineralogical studies of TG-2 core rocks suggest that secondary mineralization of permeable fractures has “sealed” the outflow area, restricting flow. Analysis of the temperature profiles and flow behavior of TG-2 suggests that the produced fluid of 359 °F (182 °C) was “pulled” into the system from elsewhere. The subsurface source of the 359 °F (182 °C) fluid is unknown and therefore targeting this resource is highly risky. Additionally, there is a high risk that exploitation of the shallow reservoir could result in rapid enthalpy declines during exploitation. The risk arises from any of the following: recharge of the reservoir by sea water; cold meteoric water influx from near surface aquifers; and breakthrough from injection wells. Groundwater influx probably poses the most significant risk. There are also a large number of connection points between the shallow thermal aquifer and the surface along the outflow path. Any pressure decline as a result of exploitation would likely allow these colder waters to descend into the reservoir and cool the production wells. Conclusions The Akutan geothermal resource can be divided into an upflow zone and one or more outflow zones. Studies of alteration minerals in the core suggest that the outflow resource discovered by TG-2 is likely to have significant permeability limitations (Stelling and Kent, 2011). The outflow resource of 359 °F (182 °C) discovered by slimhole exploratory drilling in 2010 appears to have migrated from a more distal source and may not be commercially developable. A temperature reversal at the bottom of the stabilized TG-2 profile reduces the possibility that a hotter or more voluminous reservoir would be encountered by drilling deeper at that location. These conclusions indicate that earlier estimates of production capacity of the outflow resource discovered by slimholes are inaccurate, because the flow assumptions for this estimate appear to have been overly optimistic. The hottest modern zone in the TG-2 core is at 585-590’ (178-180 m), with a static measured temperature of 338 °F (165 °C). The occurrence of wairakite, epidote and prehnite suggest that this zone was permeable during the earlier higher temperature alteration event. The outflow zone penetrated by the exploratory slimholes shows evidence of “self-sealing” through mineralization of primary and secondary permeability channels. The chlorite- and zeolite-dominated hydrothermal mineralogy of wells TG-2 and TG-4 indicate that a lower temperature alteration assemblage has been overprinted on a higher temperature assemblage. The higher temperature alteration assemblage contains illite, epidote, prehnite, and adularia. The mixed layer clays, illite-smectite and chlorite- Akutan Geothermal Resource Assessment 31 smectite, and zeolites are part of the lower temperature retrograde assemblage corresponding to temperatures of 300 – 430 °F (150-220 °C). The overprinting observations can be explained by a vertically-limited outflow system in a waning phase after attaining higher temperatures, which suggests that a deeper well drilled at or near the location of TG-2 is not likely to encounter the hotter fluids predicted by chemical geothermometry. Since those fluids appear to be flowing laterally along the water table towards Hot Springs Bay, the source fluids are probably westward up the valley towards the fumaroles and/or southwest to the valley junction and then northwest (Figs 10 and 12). Future drilling should target the upflow resource as the highest-grade, lowest-risk part of the system. The upflow source fluids are likely to be within the range of 428-572 °F (220-300 °C), and are chemically benign. The estimated output capacity of the upflow target is 15-100 MW by analog analysis, with a minimum output of 8 MW based on pessimistic volume considerations. If developing the upflow resource is not possible, the hottest part of the lower-grade outflow zone could be targeted but with greater risk. Recommendations The unresolved resource properties and risks can only be addressed by additional characterization of the resource through drilling. An evaluation of the access, drilling and financial issues involved in targeting a well on the upflow resource below the fumarole should be top prioirity at this time. Future well(s) should be directed west beneath the fumaroles or north to a postulated upflow beneath a local resistivity high. The well should attain a minimum depth of 4500’ (1350m) to insure that it penetrates through the reservoir top, although a target depth of 6000’ (1800m) would be preferable in order to better establish reservoir thickness. As evaluation of the resource potential continues, obtaining samples of separated steam and water from the production wells will be valuable for further assessing the geochemical risks related to scaling and cold water influx. Pressure-interference and possibly tracer testing will need to be conducted as additional wells are drilled. The risks associated with drilling the upflow target could be mitigated by additional exploration work. One focus of additional exploration work could be the identification of controlling structures (likely faults). Since hot fluids are constantly plugging up the "plumbing" channels by depositing minerals in open fractures, large-scale activity on faults is required to keep the system permeable. These faults must exist, but none have been conclusively identified in HSBV or near the fumaroles. Hence, mapping large- scale active structures controlling permeability in the Akutan geothermal field could reduce well targeting risks. Aerial photography survey planned for summer 2011 may provide useful data. Additional studies (LIDAR, seismic, or other geophysical methods) could supplement this investigation after the initial review of the aerial photography. Another exploration activity that could mitigate the upflow target drilling risk is extending coverage of the MT survey further to the north and west of HSBV. This could be done with very limited additional stations (possibly 10-20) with or without a helicopter, limited to the area in the direct vicinity of the fumaroles. However, this additional data may not have a significant effect on the project risk assessment of the outflow, especially along the north rim of the outflow. Because it is likely to be relatively expensive and prone to severe wind noise, this activity is not top priority at this time. Akutan Geothermal Resource Assessment 32 If the shallow outflow is deemed to be more suitable for geothermal development, an important risk mitigation measure would be analog studies of more shallow outflows that have been developed for power generation and/or space heating. Analog studies of similar geologic environments in Iceland could be particularly useful for assessing risks associated with cold water influx and injection breakthrough. Additionally, several low-cost additional studies could help characterize the outflow resource. These include: (1) Fluid inclusion analysis of hydrothermal minerals in deposited in fractures in core rocks; (2) Analysis of the fracture orientation within well cores; (3) Detailed clay analysis identifying the percentages of illite in illite-smectite and chlorite-smectite. As with the upflow target, pressure- interference and possibly tracer testing should be conducted as additional wells are drilled. References and Bibliography Black, Robert F., 1975. Late quaternary geomorphic processes: Effects on the ancient Aleuts of Umnak Island in the Aleutians. Arctic, v. 28, p. 159-169 Coombs, D.S., Ellis, A.J., Fyfe, W.S., and Taylor, A.M. (1959) The zeolite facies, with comments on the interpretation of hydrothermal syntheses. Geoch. Cosmoch. Acta 17, 53-107. Cumming, W., 2009. Geothermal resource conceptual models using surface exploration data. Proceedings, 34th Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford, CA. Cumming, W. and Mackie, R., 2010. Resistivity Imaging of Geothermal Resources Using 1D, 2D and 3D MT Inversion and TDEM Static Shift Correction Illustrated by a Glass Mountain Case History. Proceedings, World Geothermal Congress 2010. Garg and Combs, 2010. A Reexamination of USGS Volumetric “Heat in Place” Method. Proceedings, Thirty-Sixth Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford, California, SGP-TR-191, 5p. Glassley, W., 2008. Geothermal Energy: Renewable Energy and the Environment. CRC Press, 290p. Giggenbach, W., 1991. Chemical Techniques in Geothermal Exploration. In: The Application of Geochemistry in Geothermal Reservoir Development, F. D'Amore Ed. 1991 UNITAR/UNDP Guidebook. Grant, M.A., and P.F. Bixley, 2011. Geothermal Reservoir Engineering, 2nd Edition. El Sevier Press, 378 p. Information Insights, 2010. Akutan Geothermal Energy Demand and Stakeholder Assessment. Unpublished report to the City of Akutan and the Alaska Energy Authority, 34p. Kolker, A., 2008. Geologic Setting of the Central Alaskan Hot Springs Belt: Implications for Geothermal Resource Capacity and Sustainable Energy Production. Ph.D. Dissertation, University of Alaska Fairbanks, 203p. Available Online at: http://www.uaf.edu/rap/students/Alumni/Kolker-dissertation-2008.pdf Kolker, A., and R. Mann, 2009. Heating Up The Economy With Geothermal Energy: A Multi-Component Sustainable Development Project at Akutan, Alaska. Geothermal Resource Council Transactions No. 33, 11p. Kolker, A., P. Stelling, B. Cumming, A. Prakash, and C. Kleinholt, 2009. Akutan Geothermal Project: Report on 2009 Exploration Activities. Unpublished report to the City of Akutan and the Alaska Energy Authority, 37p. Kolker, A., B. Cumming, and P. Stelling, 2010. Akutan Geothermal Project: Preliminary Technical Feasiblity Report. Unpublished report to the City of Akutan and the Alaska Energy Authority, 31p. Akutan Geothermal Resource Assessment 33 Kolker, A., B. Cumming, and P. Stelling, 2010. Geothermal Exploration at Akutan, Alaska: Favorable Indications for a High-Enthalpy Hydrothermal Resource Near a Remote Market. Geothermal Resource Council Transactions No. 34, 14p. Kolker, A., A. Bailey, W.T. Howard, 2011. Akutan Geothermal Project: Summary of Findings from the 2010 Exploratory Drilling Program. Unpublished report to the City of Akutan and the Alaska Energy Authority, 33p. Lu, Z., C. Wicks, D. Dzurisin, W. Thatcher, and J. Power, 2000. Ground Deformation Associated with the March 1996 Earthquake swarm at Akutan Volcano, Revealed by Satellite Radar Interferometry. Journal of Geophysical Research, v. 105, No. B9, p. 21483-21495. Miller, T.P., G. McGimsey, D. Richter, J. Riehle, C. Nye, M. Yount, and J. Dumoulin, 1998. Catalog of the historically active volcanoes of Alaska. USGS Open-file Report 98-582. Motyka, R., and C. Nye, eds., 1988. A geological, geochemical, and geophysical survey of the geothermal resources at Hot Springs Bay Valley, Akutan Island, Alaska. Alaska Division of Geological and Geophysical Surveys (ADGGS), Report of Investigations 88-3. Motyka, R.J., S. Liss, C. Nye, and M. Moorman, 1993. Geothermal Resources of the Aleutian Arc. ADGGS Professional Paper 114. Newhall, C.G., and D. Dzurisin, 1988. Historical unrest at large calderas of the world. USGS Bulletin 1855. Newendorp, P. and Schuyler, J., 2000. Decision Analysis for Petroleum Exploration, Second Edition. Planning Press, pp. 618. Powell, T., and W. Cumming., 2010. Spreadsheets for Water and Geothermal Gas Chemistry. Proceedings of the Thirty-Fifth Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford, California, SGP-TR- 188. Richter, D.H., C.F. Waythomas, R.G. McGimsey, and P.L. Stelling, 1998. Geology of Akutan Island, Alaska. USGS Open-File Report 98-135, 1 sheet, 1:63,360 scale Rohrs, D., 2011. “Geochemistry of the Akutan Geothermal Prospect, Alaska.” Unpublished report to City of Akutan, 36p. Seki, Y., Onuki, H., Okumura, K., and Takashima, I. (1969b) Zeolite distribution in the Katayama geothermal area, Onikobe, Japan. Jap. J. Geol. Geogr. 40, 63-79. Simkin, T., and L. Siebert, 1994, Volcanoes of the World, 2nd edition: Geoscience Press in association with the Smithsonian Institution Global Volcanism Program, Tucson AZ, 368 p. Symonds, R. B., R. Poreda, W. C. Evans, C. J. Janik, and B. E. Ritchie, 2003. Mantle and crustal sources of carbon, nitrogen, and noble gases in Cascade-Range and Aleutian-Arc volcanic gases. USGS Open-File Report 03-436. Waythomas, C.F, J.A. Power, D.H. Richter, and R.G. McGimsey, 1998. Preliminary Volcano-Hazard Assessment for Akutan Volcano, East-Central Aleutian Islands, Alaska: U.S. Geological Survey Open-File Report OF 98-0360, 36 p., 1 plate, scale unknown. Western Governors’ Association Clean and Diversified Energy Initiative Report, 2006. http://www.westgov.org/wga/initiatives/cdeac/Geothermal-full.pdf Akutan Geothermal Resource Assessment 34 WesternGeco, 2009. Magnetotelluric Survey at Hot Springs Bay Valley, Akutan, Alaska: Final Report – 3D Resistivity Inversion Modeling. Unpublished report prepared for the City of Akutan, Alaska, GEOSYSTEM/WesternGeco EM, Milan, Italy, 27p. Wood, C.P, 1994. Mineralogy at the magma-hydrothermal system interface in andesite volcanoes, New Zealand. Geology, 22, 75-78.