DRILLING CONFIRMATION OF THE CASITA INDICATED GEOTHERMAL RESOURCE, NICARAGUA

Similar documents
RECENT RESULTS FROM THE SAN JACINTO - TIZATE GEOTHERMAL FIELD, NICARAGUA

HIGH TEMPERATURE HYDROTHERMAL ALTERATION IN ACTIVE GEOTHERMAL SYSTEMS A CASE STUDY OF OLKARIA DOMES

SUB-SURFACE GEOLOGY AND HYDROTHERMAL ALTERATION OF WELLS LA-9D AND LA-10D OF ALUTO LANGANO GEOTHERMAL FIELD, ETHIOPIA

WAMUNYU EDWARD MUREITHI I13/2358/2007

Magneto-telluric (MT) surveys in a challenging environment at Lihir Island, PNG

GeothermEx, Inc. GEOTHERMAL RESERVOIR ASSESSMENT METHODOLOGY FOR THE SCIENTIFIC OBSERVATION HOLE PROGRAM, KILAUEA EAST RIFT ZONE, HAWAII TASK 1 REPORT

OVERVIEW OF THE WAIRAKEI-TAUHARA SUBSIDENCE INVESTIGATION PROGRAM

INTERPRETATION OF INTERFERENCE EFFECTS IN THREE PRODUCTION WELLS IN THE KAWERAU GEOTHERMAL FIELD, NEW ZEALAND. Lynell Stevens and Kevin J Koorey

Structural Controls on the Chemistry and Output of the Wells in the Olkaria Geothermal Field, Kenya

Geological Evaluation of the Waringin Formation as the host of a Vapor-Dominated Geothermal Reservoir at the Wayang Windu Geothermal Field

Exploration of Geothermal High Enthalpy Resources using Magnetotellurics an Example from Chile

Determination of Calcite Scaling Potential in OW-903 and OW-914 of the Olkaria Domes field, Kenya

SUBSURFACE HYDROTHERMAL ALTERATION AT THE ULUBELU GEOTHERMAL FIELD, LAMPUNG, SOUTHERN SUMATRA, INDONESIA. Suharno 1, 2 and PRL Browne 2

Geothermal Model of the Lahendong Geothermal Field, Indonesia

GEOTHERMAL ENERGY POTENTIAL FOR LONGONOT PROSPECT, KENYA. By Mariita N. O. Kenya Electricity Generating Company

Interpretation of Subsurface Geological Structure of Massepe Geothermal Area Using Resistivity Data

NICARAGUA COUNTRY UPDATE. Ariel Zúñiga and Mauricio Medina. Instituto Nicaragüense de Energía, Apartado Postal 3226

Geothermal Exploration in Eritrea

Heat (& Mass) Transfer. conceptual models of heat transfer. large scale controls on fluid movement. distribution of vapor-saturated conditions

Structural Geology tectonics, volcanology and geothermal activity. Kristján Saemundsson ÍSOR Iceland GeoSurvey

HYDROTHERMAL ALTERATION IN THE SUNAGOHARA FORMATION, OKUAIZU GEOTHERMAL SYSTEM, JAPAN

BULLS-EYE! - SIMPLE RESISTIVITY IMAGING TO RELIABLY LOCATE THE GEOTHERMAL RESERVOIR

Geophysical Surveys of The Geothermal System of The Lakes District Rift, Ethiopia

RESPONSE OF WAIRAKEI GEOTHERMAL RESERVOIR TO 40 YEARS OF PRODUCTION

The Initial-State Geochemistry as a Baseline for Geochemical Monitoring at Ulubelu Geothermal Field, Indonesia

Taller de Geotermica en Mexico Geothermal Energy Current Technologies

Geochemical Characteristics of Reservoir Fluid from NW-Sabalan Geothermal Field, Iran

GEOTHERMAL POTENTIAL OF ST. KITTS AND NEVIS ISLANDS

Geothermal Exploration in Eritrea

Wind Mountain Project Summary Memo Feeder Program

TECHNICAL REVIEW OF GEOTHERMAL POTENTIAL OF NGOZI AND SONGWE GEOTHERMAL PROSPECTS, TANZANIA.

Savo Island Inferred Geothermal Resource Assessment

Project Highlights. Mineral rights include three mining concessions (MDs) covering a total área of ha, 100% owned by Minera Sud Argentina S.A.

The Role of Magnetotellurics in Geothermal Exploration

CHARACTERIZING FEED ZONES IN GEOTHERMAL FIELDS: INTEGRATED LEARNINGS FROM COMPLETION TESTING, IMAGE LOGS AND CONTINUOUS CORE

Update of the Geologic Model at the Las Pailas Geothermal Field to the East of Unit 1

Northern Chile, 85 Km NE of La Serena 125 km S of Relincho (Nueva Union Teck-Goldcorp)

Bog Hot Valley. (updated 2012)

Updated Resource Assessment and 3-D Geological Model of the Mita Geothermal System, Guatemala

Activity Submitted by Tim Schroeder, Bennington College,

GEOTHERMAL DEVELOPMENT IN THE COMOROS AND RESULTS OF GEOTHERMAL SURFACE EXPLORATION

Characterization of Subsurface Permeability of the Olkaria East Geothermal Field

NUMERICAL MODELING STUDY OF SIBAYAK GEOTHERMAL RESERVOIR, NORTH SUMATRA, INDONESIA

Summary of geothermal setting in Geo-Caraïbes target islands

GEOPHYSICAL STUDY OF THE NORTHERN PART OF TE KOPIA GEOTHERMAL FIELD, TAUPO VOLCANIC ZONE, NEW ZEALAND

Geochemical monitoring of the response ofgeothermal reservoirs to production load examples from Krafla, Iceland

Aliabad-Morvarid iron-apatite deposit, a Kiruna type example in Iran

From Punchbowl to Panum: Long Valley Volcanism and the Mono-Inyo Crater Chain

Cubic Spline Regularization Applied to 1D Magnetotelluric Inverse Modeling in Geothermal Areas

Topic 7: Geophysical and Remote Sensing Models. Presenter: Greg Ussher Organisation:SKM

GEOTHERMAL RESOURCE CONCEPTUAL MODELS USING SURFACE EXPLORATION DATA

TEMPERATURE GEOTHERMAL SYSTEM *.BY. Roger F. Harrison Salt Lake City, Utah. C; K. Blair

Hydrothermal Alteration of SMN-X,Sumani Geothermal Area, West Sumatra, Indonesia

Numerical Simulation of Devolution and Evolution of Steam-Water Two-Phase Zone in a Fractured Geothermal Reservoir at Ogiri, Japan

Exploration results and resource conceptual model of the Tolhuaca Geothermal Field, Chile

THREE DIMENSIONAL NUMERICAL MODELING OF MINDANAO GEOTHERMAL PRODUCTION FIELD, PHILIPPINES

Exploration results and resource conceptual model of the Tolhuaca Geothermal Field, Chile

Geothermal Systems: Geologic Origins of a Vast Energy Resource

Mita, a Newly Discovered Geothermal System in Guatemala

Integrated Geophysical Model for Suswa Geothermal Prospect using Resistivity, Seismics and Gravity Survey Data in Kenya

Three Dimensional Inversions of MT Resistivity Data to Image Geothermal Systems: Case Study, Korosi Geothermal Prospect

ARGENTINE FRONTIER RESOURCES INC (AFRI) SALTA EXPLORACIONES SA (SESA)

INTEGRATED GEOPHYSICAL STUDIES OF THE ULUBELU GEOTHERMAL FIELD, SOUTH SUMATERA, INDONESIA

INFRARED AND SATELLITE IMAGES, AERIAL PHOTOGRAPHY

POTENTIAL OF THE DONGWE COPPER GOLD PROJECT

Geo-scientific Data Integration to Evaluate Geothermal Potential Using GIS (A Case for Korosi-Chepchuk Geothermal Prospects, Kenya)

Volcanic Strata-Hosted Gold Deposits in Quaternary Volcanoes: The Sandwich-Style Model

Numerical Simulation Study of the Mori Geothermal Field, Japan

THREE DIMENSIONAL INVERSIONS OF MT RESISTIVITY DATA TO IMAGE GEOTHERMAL SYSTEMS: CASE STUDY, KOROSI GEOTHERMAL PROSPECT.

Hydrothermal Chemistry/ Reverse Weathering. Marine Chemistry Seminar

CSA Mine Observations Applied to the Development of Regional Exploration Models

PERFORMANCE EVALUATION OF DEEPENED WELLS 420DA AND 517DA IN THE LEYTE GEOTHERMAL PRODUCTION FIELD, PHILIPPINES

CALORIMETRY AND HYDROTHERMAL ERUPTIONS, WAIMANGU HYDROTHERMAL FIELD,

Ann Moulding and Tom Brikowski University of Texas at Dallas, Department of Geosciences

STRATIGRAPHY AND HYDROTHERMAL ALTERATION ENCOUNTERED BY MONITOR WELLS COMPLETED AT NGATAMARIKI AND ORAKEI KORAKO IN 2011

Kizito Maloba Opondo. Kenya Electricity Generating Company

THE FLUID CHARACTERISTICS OF THREE EXPLORATION WELLS DRILLED AT OLKARIA-DOMES FIELD, KENYA

THE ZUNIL-II GEOTHERMAL FIELD, GUATEMALA, CENTRAL AMERICA

NOTICE CONCERNING COPYRIGHT RESTRICTIONS

Keywords: geophysics field camp, low to moderate, temperature, geothermal system, Mt Pancar, Indonesia

3D MAGNETOTELLURIC SURVEY AT THE YANAIZU-NISHIYAMA GEOTHERMAL FIELD, NORTHERN JAPAN

LOS AZULES April 2018

Quaternary clays alluvial sands of the Shepparton Formation overlie the basement rocks.

Chapter 7: Volcanoes 8/18/2014. Section 1 (Volcanoes and Plate Tectonics) 8 th Grade. Ring of Fire

Heat Loss Assessment of Selected Kenyan Geothermal Prospects

GEOTHERMAL STRUCTURE AND FEATURE OF SULFIDE MINERALS OF THE MATALOKO GEOTHERMAL FIELD, FLORES ISLAND, INDONESIA

Preliminary results of deep geothermal drilling and testing on the Island of Montserrat

DIRECTIONAL WELLS AT THE PAILAS GEOTHERMAL FIELD, SUPPORTING GEOLOGICAL AND STRUCTURAL DATA FOR THE GEOTHERMAL MODEL

L wave Lahar Lava Magma

Tu D Understanding the Interplay of Fractures, Stresses & Facies in Unconventional Reservoirs - Case Study from Chad Granites

Conceptual model for non-volcanic geothermal resources - examples from Tohoku Japan

Preparation for a New Power Plant in the Hengill Geothermal Area, Iceland

Case Study: Tauhara New Zealand. Santiago de Chile, May 2014

Japan Engineering Consultants, Inc., Energy and Industrial Technology Development Organization,Tokyo, Japan

Integration of Surface and Well Data to Determine Structural Controls on Permeability at Salak (Awibengkok), Indonesia

Fletcher Junction Project Technical Update December 18, 2008

Status of geothermal energy exploration at Buranga geothermal prospect, Western Uganda

Thermal Modeling of the Mountain Home Geothermal Area

INTERGRATED GEOPHYSICAL METHODS USED TO SITE HIGH PRODUCER GEOTHERMAL WELLS

Structural Controls on the Geochemistry and Output of the Wells in the Olkaria Geothermal Field of the Kenyan Rift Valley

Transcription:

DRILLING CONFIRMATION OF THE CASITA INDICATED GEOTHERMAL RESOURCE, NICARAGUA P. White 1, K. Mackenzie 2, J. Brotheridge 2, J. Seastres 2, B. Lovelock 2 and M. Gomez 3 1 Panda Geoscience, M384 10/215 Rosedale Road, Albany, Auckland 0632, New Zealand 2 SKM Limited, PO Box 9806, Newmarket, 3 Cerro Colorado Power S.A., Ofiplaza El Retiro #723, Rotonda El Periodista, 150 mts al sur, Managua, Nicaragua pwhite@pandageoscience.co.nz Keywords: Casita, Nicaragua, slimhole, Indicated Resource ABSTRACT Following on from previous geological, geochemical and geophysical exploration activities, a vertical slimhole well (CSS-01) on the mid slopes of Casita volcano was completed to a depth of 842 m in November 2011 by Cerro Colorado Power (CCP). Wireline logging and flow testing of this well has confirmed the existence of a vapourdominated reservoir at Casita, with a temperature of 230-250 C, and a total gas content of about 0.7 wt%. Based on the results from CSS-01, MT geophysics, and the distribution of surface thermal activity, the Indicated geothermal resource within the CCP concession covers an area of up to 6 km 2 as per the definition of the Canadian Geothermal Reporting Code. There is likely to be a liquid reservoir below the vapour zone around an upflow under Casita volcano, although this has yet to be proven by drilling. Well CSS-01 (which has provided temperature and chemistry data) and the surrounding area with thermal activity constitutes an Indicated Resource. Prospective areas outside the Indicated Resource area, including La Pelona caldera and the lower flanks of Casita, are classified as Inferred Resource, based on the MT geophysical data. A resource assessment prepared by SKM for CCP produced the following estimates: The geothermal system within the Casita San Cristobal concession generally underlies and extends northeast from Volcán Casita. The active volcanic centre of Volcán San Cristobal lies immediately to the west (Figure 2). 1.2 Previous Work The geothermal potential of the Casita prospect was initially recognised following regional geothermal surveys undertaken by Texas Instruments, GeothermEx and Unocal. Between 1999 and 2005, Triton Minera S.A. commissioned SKM to perform integrated geological, geochemical and geophysical investigations into the geothermal potential of Casita. An initial field reconnaissance survey in 1999 was followed by a detailed structural assessment of Casita, and then a magneto-telluric (MT) geophysical survey to help delineate the likely extent and depth of the resource. This work enabled a comprehensive definition of the resource and allowed planning of an exploration drilling program, including specific exploration well targets, to proceed. The state of knowledge up to 2004 was reported by Bogie et al (2004). Subsequently, Nash (2011) incorporated the previous scientific data, together with new ASTER, SAR and a detailed (5 m) DEM into a GIS database. Some additional geological mapping and an aeromagnetic survey were undertaken by Ram Power between 2008 and 2011. Indicated Resource: Based on a probabilistic stored heat estimate and considering the direct measurements of the well CSS-01, the indicated resource within the CCP concession is 85 MWe (gross) for 20 years (at P90). Inferred Resource: Based on a probabilistic stored heat estimate, the Inferred Resource within the CCP concession area has a capacity of 68 MWe (gross) for 20 years (at P90). 1. INTRODUCTION 1.1 Location The Casita San Cristobal geothermal concession covers an area of 100 km 2 in northwestern Nicaragua (Figure 1), and is held by Cerro Colorado Power S.A., which is majority owned by Ram Power Corporation. It lies approximately 20 km WNW of the San Jacinto geothermal project, which is being developed by Polaris Energy Nicaragua S.A. (also owned by Ram Power), and where construction of a 72 MWe power project is nearing completion. Figure 1: Location of Casita in northwestern Nicaragua. 2. GEOLOGY The Casita ridge is made up of a series of overlapping volcanic centres that are broadly divided into three units: the San Cristobal, Casita, and La Pelona volcanics. All three units largely comprise andesitic lavas and pyroclastics, but the San Cristobal volcanics (youngest, to the west) may contain more basaltic rocks and the La Pelona volcanics (oldest, to the east) may include an uppermost dacitic pumice layer produced by the eruption

that accompanied formation of the La Pelona Caldera. These units are considered to sit unconformably on the Upper Miocene Tamarindo Formation, described by Saginor et al. (2009) as...basaltic to andesitic lavas interlayered with thick ignimbrite deposits and volcaniclastic sediments.... Structurally, western Nicaragua is characterised by dextral trans-extension with north-south shortening (Weinberg 1992). This has produced north-west and north-east trending strike slip faults, with some associated pull-apart basins with north-south striking normal faults. This deformation has been superimposed upon a Pliocene structural regime which includes NNW and ENE strike-slip faults and north-west striking normal faults, some of which could be reactivated under the current tectonic regime. NNE and NNW striking faults predominate at Casita (Figure 2). A structural study (SKM 2002) established potential targets for exploration drilling and assisted in planning the geophysical survey. As at San Jacinto, the N to NNE striking faults are considered likely to be the most permeable structures. There is extensive argillic alteration with minor pyrite ± jarosite exposed to the east of Volcán Casita above about 900 m elevation. Some of the alteration is associated with current thermal activity, but much is not. There is a marked lack of silica deposition except for minor recrystallised residual silica crust near some of the steaming vents, consistent with a system that has a deep water level and is probably well capped. There is also a lack of sulphur, even close to the most active vents. Apart from some minor crystalline encrustations near the most active vents, alteration is principally produced by destructive steam heated processes, rather than by mineral deposition. seasonal rainfall) with temperatures up to 85 C occur in the lowlands 15 km northeast of Casita at Monte Largo and 5 km further out at El Bonete. In the intervening area there are numerous shallow wells with temperatures above ambient, although generally less than 50 C. Fumarole steam within the Ollade crater fumaroles has a gas content of 2-3 wt%, comprising 95-97% CO 2 with gas proportions typical of benign, andesitic-hosted geothermal systems. Gas geothermometry indicates source temperatures in excess of 250 C with vapour-dominated conditions, at least at shallow depths. The oxygenhydrogen isotopic composition (δ 2 H, δ 18 O) shows a δ 18 O shift of about +3.5 from local groundwater. Steaming ground on the eastern flanks of the caldera has a high proportion of entrained air, which is a natural part of these diffuse vent discharges. The isotopic composition of this steam suggests that it is derived from secondary boiling of a steam-heated aquifer. There is no direct evidence of a chloride reservoir at accessible depths beneath Casita, or in the lowlands to the northeast. If the thermal waters at Monte Largo, El Bonete (>80 C) or Las Grietas are derived from Casita, they are more likely to be outflows of steam-heated groundwater, rather than deep chloride water. If there is a deep chloride reservoir, it is likely to be saline, given the significant δ 18 O isotopic shift. 4. GEOPHYSICS A 70 station MT TDEM survey in 2004-2005 revealed a conductive layer that drapes across the system, and which was interpreted to represent a hydrothermal clay cap above the main high temperature reservoir (SKM 2005). The top of the conductor follows the terrain to some extent and rises to its highest elevation under Ollade crater and Casita ridge (Figure 3). The doming of the top of the conductor under Casita matches quite well with the area of thermal ground mapped in the area. The base of the conductor rises to its highest elevation on the SE side of Casita ridge (Figure 3). This high elevation of the base of conductor is seen to continue west beneath Ollade crater, and to the NE through the northern part of the Argentina area. A local high in the base of conductor extends through the La Pelona caldera (Figure 3). This could represent an easterly extension of the Casita system or a separate geothermal upflow zone in this area. Figure 2: Casita San Cristobal volcanic complex and geothermal concession, showing major structures (SKM 2012) 3. GEOCHEMISTRY Thermal activity consists almost entirely of steam vents and steaming ground, spread over an area of about 5 km 2 (Figure 4). The most active fumaroles are on the northern edge of Ollade crater (elevation ~1200 m). Elsewhere on the mountain, activity is more diffuse but widespread. There are no known hot springs on the mountain, although high flowrate springs (up to 15 l/s and independent of Figure 3: NW to SE 1D MT section through the NE flanks of Casita. Well CSS01 is located near point C62 near the centre of this figure, Pad C is located near point C32. The red (conductive) zone corresponds to the clay cap above the geothermal reservoir. The blue (resistive) zone corresponds to relatively fresh lavas near surface

5. WELL CSS-01 CSS-01 was drilled as a vertical well by a track mounted Christenson CS-1000 drilling rig to a total depth of 842 m between June and November 2011. Many days were spent cementing in the upper portion of the well due to repeated total losses of circulation encountered in highly permeable scoria, fractured lava flows and tuff layers. Once these zones were behind the intermediate casing, drilling down to the eventual total depth (TD) proceeded in a relatively straightforward manner, albeit with total losses of drill fluids from below the 3½ production casing shoe (PCS) at 495 m. Below 495 m, NQ-diameter (2.98 ) corehole was drilled to a depth of 842.85 m, however the NQ drill pipe became stuck in the hole at 733 m. It was decided to run a pressuretemperature (PT) wireline survey within the NQ pipe to this depth before fishing operations commenced. The PT survey was successful, however most of the pipe remained in the hole along with several other fish that were lost during the fishing operations. No further PT surveys were possible, but the well was successfully discharged. and calcite or cristobalite. At greater depth, alteration is moderate to intense, and the alteration mineralogy is dominated by one or more of illite-smectite, stilbite, and quartz. Other secondary minerals observed in these samples include calcite, chlorite or chlorite-smectite, adularia, anhydrite, kaolinite, opaques, iron oxides and chalcedony. Interlayered clays persist to TD, with the deepest sample analysed by XRD (710-713 m) containing interlayered illite-smectite with 60% smectite interlayers. The alteration minerals appear to have formed during a previous stage of the geothermal system, probably before the steam zone formed, as they indicate cooler conditions than measured temperatures and many of them indicate liquid dominated conditions. 5.3 Fluid inclusions Fluid inclusion homogenisation temperatures were measured in vuggy quartz and calcite crystals in a sample from 799.4 m. These yielded temperatures of 250 C for quartz and 263 C for calcite. These temperatures are consistent with boiling point for depth and are slightly hotter than measured downhole temperatures. They are also out of equilibrium within the mineral geothermometry, which suggests a complex thermal history. 5.4 Permeability Multiple lost circulation zones were encountered throughout the upper part of this well, where coarse scoriaceous material was highly permeable, and the many alternating layers within the lavas meant that no sooner had one zone been plugged than another loss zone was drilled into. Numerous cement plugs were used to plug these losses and regain circulation, and were a major factor in the length of time required to drill the well. Figure 4: Location of slimhole well CSS01, along with other potential well pad sites, major structures (blue lines), thermal features (red dots), hydrothermal alteration (pink shading), access roads, and the concession boundary 5.1 Geology The lithologies encountered in CSS-01 comprised predominantly andesitic to basaltic lava flows with minor intercalated tuff, lithic tuff and breccias, and some thin smectite-rich tuffaceous and scoriaceous material. No ignimbrites or pumice of the La Pelona Formation were encountered, indicating that the La Pelona caldera may not extend as far under the Casita volcanic pile as originally thought. Similarly, the Tamarindo Formation was not distinguished in this well. 5.2 Alteration Alteration is weak in the top 250 m of CSS-01, and dominated by smectite, interlayered illite-smectite, kaolinite and iron oxides. The lavas are less altered than the tuffs, with only minor clay, iron oxides and possibly zeolites, mostly confined to the groundmass. Between 250 and 400 m, alteration is moderately intense, dominated by illite-smectite clay, with minor iron oxides There were no specific structural targets for this well and being vertical it was not likely to intersect any near vertical structures. However at 503 m, drilling encountered a void space of 1.3 m (i.e. no core was recovered). The formation above and below this void appears to be the same, so the permeability was not related to a contact between two different lithologies. Below the void, the andesite was brecciated and cemented by quartz and calcite, suggesting that it might represent a major fault/permeable zone. 5.5 Temperatures The well prognosis estimated that the top of the reservoir would be encountered at approximately 500 m based on MT data. A maximum reading thermometer (MRT) survey was carried out during a break in the drilling at 544 m, and periodically after that down to 839 m. Generally the well was allowed to heat-up for 30 minutes or more, and then the thermometer was left at bottom hole for another 30 minutes. These measurements showed a rapid temperature increase from about 160 to 230ºC around 600 m, where possibly the top of a steam zone was encountered, and then temperatures were more or less isothermal to TD. The maximum temperature recorded was 232 C at a 696 m. The depth of the water table was determined to be at 275 m when the downhole PT survey was run. On 25th October 2011, completion testing was carried out on CSS-01 behind NQ pipe down to 724 m. This included a shut-in survey prior to an injectivity test and a short

pressure fall-off (PFO) test. The maximum temperature recorded at maximum logged depth was 225ºC (Figure 5). increase near the maximum logged depth around 710 m, but this could be condensate at the top of the fish. The immediate shut survey 2 hours after injection showed recovery of downhole temperatures inside casing, but temperatures near the maximum logged depth were less than during injection. An estimate of the injectivity index from downhole pressures near the PCS is around 1.4 t/hr/bar, consistent with the total loss of circulation during drilling. The 2 hour PFO data indicates a good estimated permeability-thickness of around 0.9 darcy-meter. 5.6 Discharge testing The well was initially shut-in with a standing wellhead pressure around 100 psig (6 barg) and then gradually opened to reach 100% opening at around 10:00am on 6 December 2011. The well discharge initially produced some hot water but quickly became a dry steam discharge after only a couple of minutes. Monitoring of the well s physical parameters commenced immediately at regular intervals over the following 10 days. Figure 5: Downhole temperature profiles from CSS01 measured during injectivity tests Wellbore modelling indicates that at a typical commercial wellhead pressure of 5 barg, approximately 4 tons/hr of steam flow could be produced, even with the restrictions within the wellbore due to the fish left in the hole. This steam flow equates to a power output of about 0.5 MWe assuming a steam conversion rate of 7.4 tons/hr-mwe. The discharge steam from CSS-01 has a low gas content of 0.7 wt%, comprising mainly CO 2 (75% by volume) with relatively high H 2 S (22%). The low gas content and isotopic composition (close to local groundwater) contrast strongly with the Ollade crater fumarole chemistry and suggest the CSS-01 steam is derived from late-stage boiling of a diluted reservoir water. There is considerable variability in gas geothermometer temperatures, but overall the gas proportions and geothermometry indicate deep source temperatures in excess of 250 C. The discharge gas chemistry was stable over the 7 days of sampling, suggesting that it is representative of the reservoir steam. Figure 6: Downhole pressure profiles from CSS-01 measured during injectivity tests The injection test at 180 l/min for 2 hours produced a decline in temperatures from surface to the production casing shoe (PCS) at 494 m. An increase in temperature after the PCS suggests that most of the injected fluid was lost just below the PCS. Interestingly, downhole temperatures down to maximum logged depth (225 C) are higher than the shut temperature profile prior to the injection test. None of these temperatures were quite as high as those indicated by the MRT during drilling. The water level during the injection test was at approximately 275 m (Figure 6). A change in pressure gradient from ~508 m to ~710 m suggests the presence of less dense fluid (vapour phase). There is a slight pressure Figure 7: Discharge testing of well CSS-01 7. RESOURCE ASSESSMENT The Casita geothermal system was assessed in terms of a geothermal Resource, based on the reporting requirements specified by the Canadian Geothermal Code (CanGEA, 2010). Geothermal Reserves estimates, based on a full

commercial and environmental assessment have not been considered. The resource estimates have been made using probabilistic stored heat calculations, assuming a base temperature of 50ºC and a cut-off temperature of 180ºC. The other assumptions behind these estimates are defined in a report to Cerro Colorado Power (SKM 2012). The area around well CSS-01 (which gives temperature and chemistry data) and the surrounding area with thermal activity constitutes the Indicated Resource. Confidence in the Indicated Resource has been increased by the drilling and flow testing of well CSS-01. Prospective areas outside the Indicated Resource area, including La Pelona and the lower flanks of Casita, are classified as Inferred Resource on the basis of other evidence, in particular MT geophysics and surface thermal manifestations. Based on probabilistic stored heat estimates and considering the direct measurements of well CSS-01, the Indicated Resource within the CCP concession is 85 MWe (gross) for 20 years (at P90). The total stored heat corresponds to about 2240 PJ (P90). Based on probabilistic stored heat estimates, the Inferred Resource within the CCP concession area has a capacity of 68 MWe (gross) for 20 years (at P90). The total stored heat corresponds to about 1900 PJ (P90) relative to 50 C. The Inferred Resource does not include the area of the Indicated Resource as they have different levels of confidence and thus different parameters are used in the calculations. The Indicated Resource has a much higher level of confidence than the Inferred Resource, and within the capacity ranges presented, the P90 value is the most conservative. A gross conversion factor of 17% has been utilised. This is based on converting the recovered heat into electricity, which is largely a function of reservoir temperature, though several other factors such as silica saturation, separator pressures etc. are also important. It is assumed the power plant will be a condensing steam turbine, and a steam consumption rate of 7.4 t/hr/mwe has been used (similar to the new Fuji units at San Jacinto). 8. CONCLUSION Completion of a slimhole well at Casita has enabled the temperature, chemistry and state of the geothermal reservoir to be characterised, and provided additional confidence in the resource estimates. Cerro Colorado Power is now in the process of obtaining exploitation rights to the geothermal resource, with the objective of mobilising a production drilling rig and proceeding with development drilling. ACKNOWLEDGEMENTS Cerro Colorado Power S.A. and Ram Power Corp. are thanked for their permission to publish this work. REFERENCES Bogie, I.; Ussher, G.N.; Lawless, J.V.: The Casita geothermal field, Nicaragua. Proc. 26 th New Zealand Geothermal Workshop, Taupo, New Zealand (2004). CanGEA.: The Canadian Geothermal Reporting Code for Public Reporting, first edition. Published by the Canadian Geothermal Energy Association (2010), (www.cangea.ca) Nash, G.D.: Volcán Casita, Nicaragua: a GIS Based geothermal assessment. Proc. GRC Transactions 35, (2011). Saginor, I.; Gazel, E.; Carr, M.J.: Mid Miocene volcanism in Nicaragua and implications for the formation of the Nicaraguan Depression. American Geophysical Union, Fall Meeting 2009, abstract #V51C-1691 (2009). SKM: Structural study of the Casita geothermal field. Report for SAJENCO (2002). SKM: Casita geothermal prospect scientific assessment. Unpublished report for Polaris Energy Nicaragua (2005). SKM: Casita-San Cristobal geothermal project updated resource assessment. Unpublished report for Cerro Colorado Power S.A. (2012).