Loading...
HomeMy WebLinkAboutKawerak Pilgrim Hot Springs REF App_2Appendix of Attachments Kawerak REF Proposal “Pilgrim Hot Springs Geothermal Power Plant Conceptual Design” Table of Contents Detailed Scope of Work for Kawerak REF……………………………………………………….1 Scope of Work & Quote from ACEP……………………………………………………………..6 Scope of Work & Quote from CRW Engineering…………………….…………………………14 Scope of Work from Deerstone Consulting for TEDC Grant Project (pending match)…………23 Map of Kawerak Region; List of Kawerak Tribes……………………………………………….24 Kawerak Organizational Chart……………………………………………..……………………25 Map of Unaatuq-MINC-BSNC property boundaries………………………………………….…26 Pages from Final Report, Pilgrim Hot Springs Geothermal Exploration (ACEP, 2010-2014)….27 “Pilgrim Hot Springs Geothermal Power Plant Conceptual Design” Kawerak, Inc. Detailed Scope of Work 1 Kawerak, Inc. requests $372,233 from Alaska Energy Authority’s Renewable Energy Fund in order to complete the conceptual design of a geothermal power plant for Pilgrim Hot Springs, including both energy and heat production. Kawerak plans to use grant funds for project management personnel, materials and supplies, and contractual services. Alaska Center for Energy and Power (ACEP) will be the main contractor for this project, based on their experience with geothermal systems and their history working at Pilgrim Hot Springs. Between 2010 and 2014, ACEP led an extensive geothermal exploration effort at Pilgrim Hot Springs, located centrally on the Seward Peninsula. During this time period, a variety of geophysical surveys were conducted in conjunction with drilling efforts that took place during the summers of 2011, 2012, and 2013. The efforts culminated in the drilling of a large diameter well (PS 13-1) capable of high flow rates in the fall of 2013. PS 13-1 is a production well completed to a depth of 314 m with a diameter of 6 in. Natural artesian flow rates from this well are 60 gpm, while airlift-assisted flow testing of this well conducted by ACEP in September 2014 demonstrated that the well was capable of producing 300 gpm at temperatures ranging from 78.25 °C (172.85 °F) to 79.3 °C (174.74 °F). The minimal pressure changes that were measured at that time, combined with an estimated natural state heat flow from the geothermal system to be approximately 20 MWth based on thermal data collected from ground based and aerial investigations of the site, led researchers to conclude that the well has the ability to sustainably provide thermal fluid for on-site power generation and district heating applications. This well could be used as a production well to support on-site geothermal power generation using an Organic Rankine System, similar to the system installed at Chena Hot Springs. As part of this project, ACEP will work with Kawerak, Inc to hire a full-service engineering and design firm based in Alaska to support this project. This firm will be responsible for identifying project permitting requirements, reviewing geotechnical site considerations, civil design, all engineering drawings related to the conceptual design, particularly the site layout, and working with ACEP to design a cooling system and injection or discharge strategy for spend fluids consistent with state and federal regulations for wetlands areas. ACEP will support Kawerak in developing language for a Request for Proposals (RFP) for a permitting, geotechnical, and civil design subcontractor. A quote and scope description from CRW Engineering of Anchorage, Alaska is provided with this proposal for reference. As part of this effort, matching funds will be provided by Kawerak in the form of salary and materials match for project staff, as well as other match from a current Department of Energy- Office of Indian Energy Technical Assistance Grant that retains ACEP as a contractor of Kawerak. Kawerak has partnered with the Native Village of Mary’s Igloo to apply for the Bureau of Indian Affairs Division of Energy and Mineral Development – Indian Energy and Economic Development division’s Tribal Energy Development Capacity (TEDC) grant program. If the TEDC grant is awarded to Mary’s Igloo, with Kawerak as a sub-grantee, the work to be performed by Kawerak and sub-contractor Deerstone Consulting will contribute as match to the overall effort of this project and fit seamlessly into tasks related to the economic feasibility of producing geothermal power on the site. 1 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal “Pilgrim Hot Springs Geothermal Power Plant Conceptual Design” Kawerak, Inc. Detailed Scope of Work 2 The scope of work for this project will consist of: Task 1: Project Scoping and Contractor Selection Kawerak is planning to hire a full-service engineering and design firm based in Alaska to support this project. This firm will be responsible for identifying project permitting requirements, reviewing geotechnical site considerations, civil design, all engineering drawings related to the conceptual design, particularly the site layout, and working with ACEP to design a cooling system and injection or discharge strategy for spend fluids consistent with state and federal regulations for wetlands areas. ACEP will support Kawerak in developing language for an RFP for a permitting, geotechnical, and civil design subcontractor. Task 2: Design Specifications for Geothermal Power Plant and Review of Potential Equipment Vendors ACEP will develop specifications for a geothermal power plant for the Pilgrim Hot Springs site based on known characteristics of the resource combined with expected current and future site loads. Based on the data from the flow test conducted in 2014, the resource should be capable of producing 300 kW from the existing production well PS 13-1, or up to 5 MW based on the overall natural state heat flow from the system to the surrounding environment. Developing a smaller plant as a phase 1 project is an excellent way to stress test the resource over time, collecting data to both better quantify the precise upflow zone of the geothermal fluids and better assessing the maximum sustainable generation that the resource could be capable of supporting. In addition to the capacity of the resource to support power generation, it will be important to design a power plant that can operate independently of an existing grid. Many ORC systems are designed as synchronous machines and cannot operate in the absence of an existing grid. This has been a significant challenge for Chena Hot Springs, which has not been able to operate independently from a diesel system forming the backbone of the electric grid and providing necessary parameters (voltage and frequency) for the ORC to follow. This can be avoided through careful selection of the generator, or by pairing the geothermal plant with a battery/inverter system which could also help balance load and allow other energy resources to support site demand. ACEP will explore both pathways, with the goal of developing specifications for a Request for Information (RFI) soliciting vendor information for prospective equipment manufacturers. Sub-tasks in this section include load forecasting, design parameters for power plant, and equipment vendor assessment. Task 3: Support Site Infrastructure Design ACEP will work closely with engineering subcontractors retained by Kawerak, on site design to support power generation from the geothermal resource while providing for long-term sustainability of the resource. ACEP will be responsible for sharing data related to the resource on behalf of Kawerak, and communicating specifications related to the power plant design. ACEP will also support Kawerak in the review of draft plans, drawings, and infrastructure design recommendations and provide feedback as appropriate. Other tasks in this section will include identifying needed permits for future construction of the power plant, including well water withdrawal and disposal, and site development. 2 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal “Pilgrim Hot Springs Geothermal Power Plant Conceptual Design” Kawerak, Inc. Detailed Scope of Work 3 The engineering firm retained will provide a conceptual design of auxiliary power plant components, including well pump sizing and selection, piping layout from the well to the power plant and from the power plant to the discharge location, and a cooling system for the generator. They will also develop engineering design drawings, including site design and facilities layout, concept system process schematic, and provide system alternatives to consider. Task 4: Distribution System Design The Pilgrim Hot Springs geothermal site is currently considered a greenfield site, without any sort of existing permanent generation or distribution infrastructure. Site power, when needed, is currently provided by imported and portable generators. In addition to power plant design, it will be important to carefully think through the layout of the electrical distribution system. This also provides a unique opportunity to consider some non-traditional approaches that could enable Kawerak and Unaatuq, LLC to seek DOE funds. Chief among these alternatives could be an entirely DC-based microgrid. While it is understood that the REF is not intended to fund nonconforming technologies, there may be value in conducting at least a preliminary assessment of alternative grid architectures as it is expected that there could be future federal funds available for innovative design in this space. As part of the distribution design, the engineering firm will develop a plan for well water disposal, including discharge to the surface through a pool/pond, and/or discharge to groundwater through an underground injection well. The engineering firm will also identify additional data or other information that will be required for permitting efforts. Task 5: Assessment of Alternatives ACEP will conduct a due diligence analysis of alternative energy sources at the site. In addition to traditional diesel generation, ACEP will consider solar as an alternative. This aligns well with an initial summer-only operation as proposed by Kawerak. Surveys of Alaska installations show costs ranging from $2.20 to $5.00/Watt for remote installations larger than ~45 kW. Capacity factors range from 8%-16%. For the Nome area, according to NREL's PVWatts tool, solar irradiance levels are on the order of 5-6 kWh/m2/day in the summer months, and a ~300 kW array is predicted to yield a little over 300,000 kWh total annual energy production. Alaska’s cold temperatures increase system voltage, reduce electrical resistance, and yield higher-than-rated outputs associated with reflected light and albedo effects. These factors, combined with declining module prices, are making solar PV technology more economical. Solar PV arrays have been installed in all areas of the state from the southwest to the Arctic, and low sun angles and long daylight hours represent opportunities to mount panels vertically on walls as well as on the east and west sides of buildings. Task 6: Economic Feasibility Analysis Based on the information derived from Tasks 2-4 and on information from geothermal power plants of a similar size developed in other locations, ACEP will generate budget estimates and complete a robust economic analysis for the project as well as alternatives identified through Task 5 work. ACEP will work closely with Kawerak and the engineering firm to develop design and construction cost estimates. The engineering firm will provide estimates for one preferred system design. 3 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal “Pilgrim Hot Springs Geothermal Power Plant Conceptual Design” Kawerak, Inc. Detailed Scope of Work 4 The TEDC Grant, if awarded, will contribute significantly to the economic analysis portion of this project. As economic factors are determined and the conceptual design is developed, Deerstone Consulting, as part of their TEDC project scope, will work with all project participants and stakeholders, led by the Native Village of Mary’s Igloo (applicant for the TEDC grant) and Kawerak to assist in the following tasks and deliverables: • Facilitate a Strategic Planning Session with Unaatuq Board members with advisory support from the Native Village of Mary’s Igloo and the Bering Strait Regional Energy Leadership Committee to define scenarios and energy development options for Pilgrim Hot Springs and more broadly for the Kawerak/Norton Sound region. • Perform an Economic Analysis of the income potential and requirements for an energy provider as well as a load analysis for various prospective businesses and enterprises based on development scenarios identified by project participants and stakeholders. • Conduct a Feasibility Study to determine an appropriate business structure and define development options for Unaatuq, LLC to optimize and manage the locally available energy resources—including possibly geothermal, solar, biomass, hydropower, and wind—for a sustainable economy and business foundation for Pilgrim Hot Springs. The objectives for the study are broken down as follows: • Determine what type of business structure is best suited to the energy development goals of the region (profit vs. nonprofit) • Develop options for a conceptual model for selling geothermal power at Pilgrim Hot Springs to outside entities • Outline the economic benefits of developing a tribally-led business utility structure including expected income, cash flow, and regional economic benefits • Determine how the business structure could also develop regulations, best practices, and resources to other regional utilities and standalone power plants • Engage Unaatuq with the Bering Strait Regional Energy Leadership Committee as an advisory group regarding energy development strategies Task 7: Stakeholder Education As part of this project, the project team and members of the Unaatuq Board Advisory Committee will travel to Chena Hot Springs in Fairbanks, Alaska to view the current geothermal power plant system at their hot springs resort. The team will meet with Chena management and staff to discuss strategies, challenges, benefits and risks associated with the system and business. The project team and Unaatuq Board Advisory Committee members will also visit the University of Alaska Fairbanks’ Campus and the Alaska Center for Energy and Power to meet with ACEP staff in person to discuss project progress, geothermal technologies, and other appropriate microgrid technologies that may be applicable to providing energy at Pilgrim Hot Springs. Task 8: Stakeholder Engagement As part of this project, two meetings will be held on site at Pilgrim Hot Springs for all project staff members, contractors, and the Unaatuq Board of Directors. During these stakeholder engagement meetings, which will take place in conjunction with ACEP’s visits to the site, Unaatuq Board Members as well as Kawerak Project Staff will come 4 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal “Pilgrim Hot Springs Geothermal Power Plant Conceptual Design” Kawerak, Inc. Detailed Scope of Work 5 together to view possible locations of the power plant components, water piping and distribution ideas, examine the monitoring wells and production well, and discuss options for waste water disposal. By keeping the Unaatuq Board of Directors, as well as other stakeholders, in consistent contact with Kawerak Project Staff and project contractors, there will be transparency in the project details as well as valuable opportunities for feedback, suggestions, and innovative ideas suited to the goals of the board; to help them achieve their mission and vision. Unaatuq Mission: To promote the wellbeing of our people through sharing, protecting, and responsibly developing the resources of Pilgrim Hot Springs. Unaatuq Vision: A protected arctic oasis that provides for our people. Task 9: Ongoing Grant Project Management and Financial Reporting Kawerak will work with closely with project staff and contractors to keep on task with project milestones and deliverables. Reports will be provided to Alaska Energy Authority in a timely manner and as described in the grant agreement. Financials will be recorded, documented and reported on a consistent basis throughout the life of the grant. All project activities will be shared with appropriate stakeholders, such as the Kawerak Board of Directors and the Unaatuq Board of Directors, on a consistent basis in the form of email reports, verbal updates, presentations at board meetings, and through department newsletters. 5 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Office of Grants and Contracts Administration P.O. Box 757880, Fairbanks, Alaska 99775-7880 Tapiana Wray Principal Grant & Contract Management Officer (907) 474-1989 phone (907) 474-5506 fax uaf-ogca-preaward@alaska.edu tewray@alaska.edu www.uaf.edu/ogca/ UA is an AA/EO employer and educational institution and prohibits illegal discrimination against any individual: www.alaska.edu/titleIXcompliance/nondiscrimination. September 23, 2020 Kawerak, Inc. P.O. Box 948 Nome, Alaska 99762 Re: Letter of Commitment for proposal to Alaska Energy Authority 2020 Renewable Energy Fund The University of Alaska Fairbanks is pleased to collaborate with Kawerak, Inc., on the proposal entitled “Pilgrim Hot Springs Geothermal Power Plant Conceptual Design,” which is being submitted to the Alaska Energy Authority 2020 Renewable Energy Fund. The Principal Investigator from UAF is Gwen Holdmann, Director of the Alaska Center for Energy and Power. The appropriate administrative and programmatic personnel at UAF are aware of the pertinent state and federal regulations and policies, and we are prepared to enter into a subcontract with Kawerak, Inc., that ensures compliance with all such policies, should this proposal be funded. A statement of work, budget, and budget justification for this subaward are attached. If you have questions or need additional information, please feel free to contact me at uaf-ogca- preaward@alaska.edu. Sincerely, Tapiana Wray Principal Grant & Contract Management Officer 6 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal ACEP SCOPE OF WORK PILGRIM HOT SPRINGS GEOTHERMAL POWER PLANT CONCEPTUAL DESIGN September 22, 2020 ACEP Principal Investigator Gwen Holdmann Director, Alaska Center for Energy and Power Gwen.holdmann@alaska.edu Contracting POC Rosemary Madnick Executive Director, Office of Grants and Contracts University of Alaska Fairbanks rmadnick@alaska.edu 7 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal OBJECTIVE AND INTRODUCTION The Alaska Center for Energy and Power (ACEP) led an extensive geothermal exploration effort at Pilgrim Hot Springs between 2010 and 2014. During this time period, a variety of geophysical surveys were conducted in conjunction with three major drilling efforts. The drilling program culminated in the fall of 2013 with the drilling of a 6-in diameter production well (PS 13-1) capable of high flow rates. PS 13-1 was completed to a depth of 314 m. Natural artesian flow rates from this well are 60 gpm, while airlift- assisted flow testing of this well conducted by ACEP in September 2014 demonstrated that the well is capable of producing 300 gpm at temperatures ranging from 78.25 °C (172.85 °F) to 79.3 °C (174.74 °F). The minimal pressure changes that were measured at that time, combined with an estimated natural state heat flow from the geothermal system of approximately 20 MWth based on thermal data collected from ground-based and aerial investigations of the site, led researchers to conclude that the well has the ability to sustainably provide thermal fluid for on-site power generation and district heating applications. This well could be used as a production well to support on-site geothermal power generation using an Organic Rankine Cycle (ORC) system, similar to the system installed at Chena Hot Springs. Chena Hot Springs has been operating a small geothermal power plant ranging in output from approximately 200 to 400 kWe (net), using produced fluid at 74°C (165 °F). The Principal Investigator of this project, Gwen Holdmann, was the lead for both the prior geothermal exploration at Pilgrim Hot Springs and the development of the geothermal power plant at Chena Hot Springs, and is thus well suited to lead this proposed effort. This scope of work represents a natural extension of previous work, and focuses on supporting Kawerak in developing a geothermal power plant at the site to benefit the people of the region. SCOPE OF WORK Timelines provided below are in quarters from receipt of award. TASK 1: PROJECT SCOPING AND CONTRACTOR SELECTION Kawerak is planning to hire a full-service engineering and design firm based in Alaska to support this project. This firm will be responsible for identifying project permitting requirements; reviewing geotechnical site considerations, civil design, and all engineering drawings related to the conceptual design, particularly the site layout; and working with ACEP to design a cooling system and injection or discharge strategy for spent fluids, consistent with state and federal regulations for wetland areas. ACEP will support Kawerak in developing language for an RFP for a permitting, geotechnical, and civil design subcontractor. Deliverables: Draft language for RFP, outlining suggested responsibilities Effort: 10% of total man-hours estimated Timeline: TBD 8 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal TASK 2: DESIGN SPECIFICATIONS FOR GEOTHERMAL POWER PLANT AND REVIEW OF POTENTIAL EQUIPMENT VENDORS ACEP will develop specifications for a geothermal power plant for the Pilgrim Hot Springs site, based on known characteristics of the resource and expected current and future site loads. Based on data from the flow test conducted in 2014, the resource should be capable of producing 300 kW from existing production well PS 13-1, or up to 5 MW based on the overall natural state heat flow from the system to the surrounding environment. Developing a smaller plant as a Phase 1 project is an excellent way to stress test the resource over time, collecting data to both better quantify the precise upflow zone of the geothermal fluids and better assess the maximum sustainable generation capacity of the resource. It will also be important to design a power plant that can operate independently of an existing grid. Many ORC systems are designed as synchronous machines and cannot operate in the absence of an existing grid. This has been a significant challenge for Chena Hot Springs, which has not been able to operate independently from a diesel system forming the backbone of the electric grid and providing necessary parameters (voltage and frequency) for the ORC to follow. This issue can be avoided through careful selection of the generator, or by pairing the geothermal plant with a battery/inverter system that could also help balance load and allow other energy resources to support site demand. ACEP will explore both options, with the goal of developing specifications for a Request for Information (RFI) to solicit vendor information for prospective equipment manufacturers. Task 2.1 – Load Forecasting. Develop a range of long-term load projections based on expected future infrastructure needs and use patterns. Task 2.2 – Establish optimal design parameters for power plant (input/output). Based on the known characteristics of the resource and the load forecast developed in Task 2.1, ACEP will publish design specifications for the system. Task 2.2 – Equipment Vendor Assessment. ACEP will conduct a market evaluation of equipment manufacturers, and complete due diligence research into existing products and projects on behalf of Kawerak. ACEP will also release an RFI to solicit specific vendor input that will be used for more advanced project design and cost modeling. Deliverables: Design specifications and vendor report (including results of RFI) for Kawerak. Effort: 45% of total man-hours estimated Timeline: TBD TASK 3: SUPPORT SITE INFRASTRUCTURE DESIGN ACEP will work closely with engineering subcontractors retained by Kawerak for on-site design to support power generation from the geothermal resource while providing for long-term sustainability of the resource. ACEP will be responsible for sharing data related to the resource on behalf of Kawerak, and communicating specifications related to the power plant design. ACEP will also support Kawerak in reviewing draft plans, drawings, and infrastructure design recommendations and providing feedback as appropriate. 9 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Deliverables: Data package for engineering subcontractor, written feedback on preliminary design. Effort: 10% of total man-hours estimated Timeline: TBD TASK 4: DISTRIBUTION SYSTEM DESIGN The Pilgrim Hot Springs geothermal site is currently considered a greenfield site, without any sort of existing permanent generation or distribution infrastructure. Site power, when needed, is currently provided by imported and portable generators. In addition to power plant design, it will be important to carefully think through the layout of the electrical distribution system. This provides a unique opportunity to consider non-traditional approaches that could enable Kawarak and Unaatuq, LLC to seek DOE funds. Chief among these alternatives could be an entirely DC-based microgrid. While it is understood that the Renewable Energy Fund is not intended for nonconforming technologies, there may be value in conducting at least a preliminary assessment of alternative grid architectures since future federal funds could be available for innovative design in this space. Deliverables: Distribution grid layout options, provided to engineering subcontractor to incorporate into conceptual design for site layout. Effort: 10% of total man-hours estimated Timeline: TBD TASK 5: ASSESSMENT OF ALTERNATIVES ACEP will conduct a due diligence analysis of alternative energy sources at the site. In addition to traditional diesel generation, ACEP will consider solar power as an alternative. This aligns well with an initial summer-only operation as proposed by Kawerak. Surveys of Alaska installations show costs ranging from $2.20 to $5.00/Watt for remote installations larger than ~45 kW. Capacity factors range from 8%-16%. For the Nome area, according to NREL's PVWatts tool, solar irradiance levels are on the order of 5-6 kWh/m2/day in the summer months, and an ~300 kW array is predicted to yield a little over 300,000 kWh total annual energy production. Alaska’s cold temperatures increase system voltage, reduce electrical resistance, and yield higher-than- rated outputs associated with reflected light and albedo effects. These factors, combined with declining module prices, are making solar photovoltaic (PV) technology more economical. Solar PV arrays have been installed in all areas of the state, from the southwest to the Arctic, and low sun angles and long daylight hours represent opportunities to mount panels vertically on walls as well as on the east and west sides of buildings. Deliverables: Analysis of solar and diesel as alternatives. Effort: 10% of total man-hours estimated Timeline: TBD 10 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal TASK 6: ECONOMIC FEASIBILITY ANALYSIS Based on the information derived from Tasks 2-4 and on information from geothermal power plants of a similar size developed in other locations, ACEP will generate budget estimates and complete a robust economic analysis for the project, as well as alternatives identified though Task 5 work. Deliverables: Report on economic feasibility and alternatives. Effort: 15% of total man-hours estimated Timeline: TBD 11 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal BUDGET JUSTIFICATION Budget Category Total Salaries Gwen Holdmann (executive staff, $79.25), 160 hours $15,674 Christopher Pike (exempt staff, $37.16), 160 hours $7,563 Robert Bensin (exempt staff, $37.95), 80 hours $3,862 Jeremy Vandermeer (exempt staff, $33.00), 500 hours $20,988 Stephen Colt (adjunct faculty, $370.37), 40 hours $14,844 M.S. graduate student ($23), 640 summer hours $14,735 Total salaries $77,666 Fringe benefits Executive staff, 27.6% $4,326 Exempt staff, 42.5% $13,775 Adjunct faculty, 11.1% $1,648 Graduate student, 9.7% during summer months $1,429 Graduate student health insurance $734 Total fringe benefits $21,912 Travel Fairbanks, AK-Nome, AK, 2 trips, 2 people, 3 days/nights per trip $5,436 Other Costs Poster printing, etc. $500 Truck rental $900 Total Direct Costs $106,414 Modified Total Direct Costs (MTDC) $105,514 Indirect Costs, 25% of MTDC $26,379 Total Project $132,793 Salaries • Gwen Holdmann, Director of Alaska Center for Energy and Power (UAF executive staff): 0.92 month (160 hours). Holdmann will serve as the PI and project manager, and will be responsible for communicating with Kawerak, the engineering contractor, and leading the development of specifications, RFI, and design for the geothermal power plant. She has led prior resource evaluation efforts at Pilgrim Hot Springs, and was the lead engineer for the Chena Hot Springs geothermal power plant development which is similar in size and scope to the one proposed here. • Christopher Pike (exempt staff): 0.92 month (160 hours). Pike will be responsible for data management related to site layout and PS 13-1 characteristics as well as assessment of alternative resources. Pike was the project manager for the Pilgrim Hot Springs confirmation well drilling program and is knowledgeable about each of the drill holes on site. • Robert Bensin (exempt staff): 0.46 month (80 hours). Bensin will support planning for the electrical distribution system. He is intimately familiar with the site through prior employment with Bering Straits Native Corporation and previously led the development of a small seasonal organic farm at the site. Rob is a licensed electrician and electrical administrator in the State of Alaska. 12 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal • Jeremy Vandermeer (exempt staff): 2.89 months (500 hours). Vandemeer will support the conceptual design for the electric layout for the site, as well as specifications related to the power plant and/or ancillary support equipment (energy storage, etc). Vandermeer is an electric engineer who has previously modeled integration of a geothermal power plant at Pilgrim Hot Springs into the Nome grid. • Stephen Colt (adjunct faculty): 0.46 month (80 hours) to support economic analysis. • M.S. graduate student research assistant: 640 summer hours (16 weeks at 40 hours per week). The graduate student will participate in the economic analysis of various design alternatives to support engineering and site planning decisions. Per UA policy, salaries include a leave reserve, calculated on salary at a rate of 20.6% for executive staff, 24.1% for exempt staff, 0.2% for adjunct faculty, and 0.1% for students. Salaries also include a 2.5% annual escalation for executive and exempt staff). Fringe Benefits Staff benefits are negotiated annually with the Office of Naval Research. FY21 provisional fringe benefits rates include 27.6% for executive staff, 42.5% for exempt staff, 11.1% for adjunct faculty, and 9.7% for students during the summer months. A copy of the current rate agreement is available at: http://www.alaska.edu/cost-analysis/negotiation-agreements/ Health insurance is included for the graduate student research assistant during the summer, based on academic year 2020-21 rates ($734 for summer). Travel Costs are included for two trips to Nome, Alaska, for two project personnel. Per person costs for each 3- day/3-night trip include $400 for airfare, $185/night for lodging, $118 per day for meals and incidental expenses, and $50 for ground transportation and airport parking. Costs are based on current pricing using Google flights, Department of Defense per diem rates, and previous travel to similar destinations for other research programs. Other Costs Funds are budgeted for printing posters and reports for project dissemination ($500) and for truck rental in Nome, Alaska, during fieldwork (2 trips, 3 days/trip, $150/day). Costs are based on similar costs for other research programs. Rental vehicle costs do not incur indirect costs. Indirect Costs Facilities and administrative (F&A) costs are calculated at 25% of the Modified Total Direct Costs (MTDC), based on the current MOA between UAF and the Alaska Department of Transportation and Public Facilities. MTDC exclude equipment, capital expenditures, charges for patient care, rental costs, tuition remission, scholarships and fellowships, participant support costs and the portion of each subaward in excess of $25,000. A copy of the agreement is available at: http://www.alaska.edu/cost- analysis/negotiation-agreements/. Total Direct Costs: $106,414 MTDC: $105,514 Total Indirect Costs: $26,379 Total Project: $132,793 13 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Anchorage Office: 3940 Arctic Blvd. Suite 300, Anchorage, AK 99503 | (907) 562-3252 fax (907) 561-2273 Palmer Office: 808 S. Bailey St. Suite 104, Palmer, AK 99645 | (907) 707-1352 Seattle Office: 100 S. King Street, #100-749, Seattle, WA 98104 (206) 350-2791 www.crweng.com September 24, 2020 Ms. Amanda Toerdal Pilgrim Hot Springs General Manager Pilgrim Hot Springs / Unaatuq, LLC c/o Kawerak, Inc 500 Seppala Drive Nome, AK 99762 Re: Pilgrim Hot Springs Engineering Services Dear Amanda, CRW Engineering Group (CRW) is pleased to provide this proposal for engineering services to assist Unaatuq, LLC and the Alaska Center for Energy and Power (ACEP) in developing the conceptual design for a new geothermal power plant at Pilgrim Hot Springs, Alaska. We understand that the scope and fee developed for this effort are for budgetary purposes and may be refined before the work begins. An existing 6-inch diameter geothermal production well has been drilled to 314 meters. Testing on the well has been completed with a sustained flow rate of 300 gpm and a water temperature of 175° F. The well will produce approximately 60 gpm in artesian conditions. The proposed power production for the new plant is between 100 and 250 kW, at an anticipated flow rate of 200 gpm. Specific scope of work items for CRW are outlined below: SCOPE OF WORK CRW will provide the following services: 1. Coordination with ACEP and Unaatuq LLC as needed to develop the project. 2. Identify Needed Permits a. Identify the permits needed for construction of the power plant, including: i. Well water withdrawal and disposal ii. Site development iii. Power plant construction b. Identify additional data or other information that will be required for permitting efforts. 3. Develop a plan for well water disposal, including: a. Discharge to the surface through a pool/pond. b. Discharge to groundwater through an underground injection well. 4. Provide conceptual design of auxiliary power plant components, including: a. Well pump sizing and selection b. Piping layout from the well to the power plant and from the power plant to the discharge location. c. Cooling system for the generator. d. Develop engineering design drawings, including: i. Site design and facilities layout ii. Concept system process schematic iii. One alternative only 5. Develop design and construction cost estimates. a. One estimate for preferred design. 14 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Pilgrim Hot Springs Conceptual Power Plant Design 9/24/20 2 of 3 DELIVERABLES The scope items above will be presented in an engineering technical memorandum. Conceptual design drawings, engineering calculations, design and construction cost estimates, and meeting summaries will be provided as appendices to the technical memo. ASSUMPTIONS The following assumptions have been made: · No site visits or field investigations will be required. · Existing mapping and survey data for the site and surrounding area will be sufficient to complete the conceptual design WORK BY OTHERS – It is assumed that ACEP will provide the following: · Coordination with property owners and the general public. · Mechanical and electrical design of the power plant. FEE The proposed Scope of Work will be performed on a time and materials, not to exceed, basis in accordance with the attached CRW terms and conditions for an amount of $90,960. Thank you for asking CRW to provide this proposal. We appreciate the opportunity to serve Unaatuq LLC and ask that you contact us at (907)562-3252 if you have any questions or comments. Sincerely, CRW Engineering Group, LLC. Matt Edge, P.E., Principal Direct: (907) 646-5623 / email: medge@crweng.com Attachment: 1) Engineering Services Fee Estimate Breakdown 2) CRW Engineering Group, LLC Terms and Conditions, 2020 15 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Pilgrim Hot Springs Conceptual Power Plant Design 9/24/20 3 of 3 ________________________________________________________________________ ACCEPTANCE OF THE PROPOSAL AND AUTHORIZATION TO PROCEED The proposed scope of services, fee, and contract conditions are acceptable to Unaatuq LLC, and CRW Engineering Group, LLC is authorized to proceed with the work. By ______________________________ Signature _____________________________* Title ____________________________ Date _________________________________ * Person with authority to commit the resources of Unaatuq LLC. 16 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Pilgrim Hot Springs Geothermal Power Plant Concept DesignFee Proposal CRW Engineering Group, LLCTask & Subtask DescriptionSenior Principal / CMJeff Stanley, PEPrincipal / PM (Civil/Environmental)Matt Edge, PESenior Engineer (Mechanical)Tracy McKeon, PERegistered Eng/SurveyorStaff Engineer IISenior Designer Admin SupportTotal LaborExpensesSubconsultantTotal Subtask$210$200$190$160$145$145$90$$$$Concept DesignTask 1 -Technical MemorandumProject Management & Coordination216444$5,380$100$5,480Identify Needed Permits11224$6,090$6,090Well Water Disposal Plan184328$8,370$50$8,420Conceptual DesignPump Sizing and Selection22416$3,740$25$3,765Piping Layout284824$7,300$25$7,325Cooling System1163224$11,850$25$11,875Engineering DrawingsSite Design1161640$11,530$100$11,630System Process Schematic1441624$7,810$50$7,860Cost Estimates28424484$13,940$25$13,965Technical Memo Preparation and Printing42484044$14,500$50$14,550Total Task 1:15984212414412412$90,510$450$0$90,960TOTAL ALL SERVICES:15984212414412412$90,510$450$0$90,960Assumptions1No site visits are required.2No electrical design included.3Unaatuq LLC will provide topographic information to form basis of site design.4Drawings will be in AutoCAD. No specifications will be provided. CRW Engineering Group, LLCPage 1 of 1Date: 9/24/202017 of 101Kawerak-Pilgrim Hot Springs-REF Proposal CRW Engineering Group, LLC Page 1 of 5 Standard Terms and Conditions (11/6/16) 2020 SCHEDULE OF CHARGES AND STANDARD TERMS and CONDITIONS Compensation to CRW Engineering Group, LLC for our professional services is based upon the conditions set forth below: SCHEDULE OF CHARGES Charges for employees are determined by the hourly rates listed below. A new schedule is issued at the beginning of each year. Unless other arrangements have been made, charges for all work will be based on the latest Schedule of Charges and General Conditions. EMPLOYEE CATEGORY Senior Principal..................................................................................$210.00 Principal .............................................................................................$200.00 Senior Engineer/Land Surveyor.........................................................$190.00 Project Engineer/Land Surveyor ........................................................$175.00 Registered Staff Engineer/Land Surveyor .........................................$160.00 Staff Engineer/Land Surveyor II (EIT/LSIT) ....................................$145.00 Staff Engineer/Land Surveyor I .........................................................$130.00 Senior Designer..................................................................................$145.00 Engineering/Surveying Technician III...............................................$130.00 Engineering/Surveying Technician II ................................................$115.00 Engineering/Surveying Technician I .................................................$ 95.00 Clerical/Administrative Support ........................................................$ 90.00 SUPPLIES AND SERVICES Direct Expenses and Supplies ............................................................Invoice + 10% Subconsultants ...................................................................................Invoice + 10% Meals (Per Diem) ...............................................................................$60.00/day In-house Expenses Xerox (8-1/2 x 11) ....................................................................$0.10/copy Xerox (11 x 17) .........................................................................$0.20/copy Color Copies (8-1/2 x 11) .........................................................$1.00/copy Mileage (Federal Rate) .............................................................$0.58/mile Bond Plots .................................................................................$1.00/square foot Mylar Plots ................................................................................$2.00/square foot 18 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal CRW Engineering Group, LLC Page 2 of 5 Standard Terms and Conditions (11/6/16) CRW ENGINEERING GROUP, LLC STANDARD TERMS AND CONDITIONS 1. INVOICES AND PAYMENT TERMS CRW Engineering Group, LLC (CRW) will submit invoices to CLIENT not more than every thirty (30) days and a final bill upon completion of Services. CLIENT shall notify CRW within ten (10) days of receipt of invoice of any dispute with the invoice. CLIENT and CRW will promptly resolve any disputed items. Payment on undisputed invoice amounts is due upon receipt of invoice by CLIENT and is past due thirty (30) days from the date of the invoice. CLIENT agrees to pay a finance charge of one and one-half percent (1-1/2%) per month, or the maximum rate allowed by law, on past due accounts. If payment remains past due sixty (60) days from the date of the invoice, then CRW shall have the right to suspend all work under this Agreement, without prejudice. CLIENT shall pay all reasonable demobilization and other suspension costs. CLIENT agrees to pay attorneys’ fees, legal costs and all other collection costs incurred by CRW in pursuit of past due payments. Our hourly rates do not include a sales tax and these will be added if they become applicable in any jurisdiction. 2. CHANGES CLIENT and CRW recognize that it may be necessary to modify the scope of Services, the schedule, and/or the cost estimate proposed in this Agreement. CRW shall notify CLIENT in a timely manner when it has reason to believe a change to the Agreement is warranted. CRW shall prepare a Change Order request outlining the changes to the scope, schedule, and/or cost of the project. CLIENT has a duty to promptly consider the Change Order request and advise CRW in a timely manner in writing on how to proceed. If after a good faith effort by CRW an agreement has not been reached with the CLIENT, then CRW shall have the right to terminate this Agreement upon written notice to the CLIENT. 3. DELAYS AND FORCE MAJEURE CLIENT shall not hold CRW responsible for damages or delays in performance caused by acts of God, acts and/or omissions of Federal, State, and local governmental authorities and regulatory agencies or other events which are beyond the reasonable control of CRW. 4. TERMINATION This Agreement may be terminated by either party upon written notice in the event of substantial failure by the other party to perform in accordance with terms hereof. Such termination shall not be effective if that substantial failure has been remedied before expiration of the period specified in the written notice, such period shall not be less than seven (7) calendar days. In the event of termination, CRW shall be paid for services performed to the termination notice date including reasonable termination expenses. CRW may complete such analyses and records as are necessary to complete their files and may also complete a report on the Services performed to the date of notice of termination or suspension. The expenses of termination or suspension shall include all direct costs of CRW in completing such analyses, records, and reports. 5. DATA AND INFORMATION CLIENT shall provide to CRW all the reports, data, studies, plans, specifications, documents, and other information which are relevant to the Services. CRW shall be entitled to rely upon the reports, data, studies, plans, specifications, documents, and other information provided by CLIENT or others in performing the Services, and CRW assumes no responsibility or liability for the accuracy or completeness of such. 19 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal CRW Engineering Group, LLC Page 3 of 5 Standard Terms and Conditions (11/6/16) 6. RIGHT OF ENTRY The CLIENT will furnish CRW with the right-of-entry on the land to conduct surveys, observations, and other explorations as required. If the CLIENT does not own the site, CLIENT must obtain permission for CRW to enter the site and perform the services. CRW will take reasonable precautions to minimize damage from use of equipment, however, it is understood that in the normal course of work some surface damage may occur, the restoration of which is not part of this Agreement. The CLIENT is responsible to provide, by map or drawing, a description of the property, its location, and the location of any buried utilities or structures. CRW shall not be liable for damage or injury arising from damage to subterranean structures (pipes, tanks, telephone lines, etc.) which are not called to CRW’s attention and correctly shown on the plans furnished by the CLIENT in connection with work performed under this Agreement. 7. CONTROL OF WORK AND JOB-SITE SAFETY CRW shall be responsible only for its activities and that of its employees and subcontractors. CRW’s Services under this Agreement are performed for the sole benefit of the CLIENT and no other entity shall have any claim against CRW because of this Agreement or the performance or nonperformance of Services hereunder. CRW will not direct, supervise, or control the work of other consultants and contractors or their subcontractors. 8. PROFESSIONAL WORK PRODUCT The Service provided by CRW is intended for one time use only. All documents, including but not limited to, reports, plans, designs, field data, field notes, laboratory test data, calculations, and estimates (the “Documents”), and all electronic media prepared by CRW are considered its professional work product. CRW retains all rights to its professional work product. Copies of Documents shall be provided to CLIENT upon written request and at CLIENT’s expense. CRW shall retain these Documents for a period of two (2) years following submission of its report, during which period they will be made available to CLIENT at all reasonable times. CLIENT acknowledges that electronic media is susceptible to unauthorized modification, deterioration, and incompatibility, and therefore CLIENT shall not rely on the accuracy of the electronic media version of CRW’s professional work product. CLIENT understands that the professional work product is not intended or represented by CRW to be suitable for reuse by any party, including, but not limited to, the CLIENT, its employees, agents, subcontractors or subsequent owners on any extension of a specific project not covered by this Agreement or on any other project, whether CLIENT’s or otherwise, without CRW’s prior written permission. CLIENT agrees that any reuse unauthorized by CRW will be at CLIENT’s sole risk and that CLIENT will defend, indemnify, and hold CRW harmless from any loss or liability resulting from the reuse, misuse, or negligent use of the professional work product. 9. STANDARD OF CARE Services performed by CRW will be conducted in a manner consistent with that level of care and skill ordinarily exercised by other members of the engineering and science professions currently practicing under similar conditions subject to the time limits and financial, physical, or any other constraints applicable to the Services. No warranty, express or implied, is made. 20 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal CRW Engineering Group, LLC Page 4 of 5 Standard Terms and Conditions (11/6/16) 10. INSURANCE AND INDEMNITY CRW maintains General Liability Insurance for bodily injury and property damage with an aggregate limit of $1,000,000 per occurrence and we will furnish certificates of such insurance upon request. Our liability to the CLIENT for bodily injury or property damage arising out of work performed for the CLIENT for which legal liability may be found to rest upon us, other than the professional errors and omissions, will be limited to our General Liability insurance coverage. CLIENT shall, at all times, defend, indemnify, and save harmless CRW and its subcontractors, consultants, agents, officers, directors, members and employees from and against all claims, damages, losses, and expenses, including but not limited to reasonable attorneys’ fees, court and arbitration costs, arising out of or resulting from the Services of CRW, inclusive of claims made by third parties, or any claims against CRW arising from the acts, errors, or omissions of CLIENT, its employees, agents, contractors, and subcontractors. Such indemnification shall not apply to the extent such claims, damages, losses, or expenses are finally determined to result from CRW’s negligence. CRW shall, at all times, indemnify and save harmless CLIENT and its officers, directors, agents, and employees from and against all claims, damages, losses, and expenses arising from personal injury, death, or damage to third-party property to the extent directly attributable to the negligent acts, errors, or omissions of CRW. 11. LIMITATION OF LIABILITY Our findings, recommendations, speci- fications, or professional opinions will be presented, within the limits prescribed by the CLIENT, after being prepared in accordance with generally accepted professional engineering practice. We make no other warranty, either express or implied. For any injury or loss on account of any error, omission, or other professional negligence, the CLIENT agrees to limit CRW and/or its professional employees' liability to the CLIENT and to all agents, contractors, and subcontractors arising out of the performance of our professional services, such that the total aggregate liability to all those named shall not exceed $50,000 or our fee, whichever is greater. In the event the CLIENT does not wish to limit our professional liability to this sum, we will waive this limitation upon the CLIENT's written request made at the time of the initial authorization, on a given project, provided that the CLIENT agrees to pay an additional 5% of our total fee or $500, whichever is greater. However, the CLIENT agrees that our maximum liability will be limited to our Professional Liability Insurance coverage. In the event the CLIENT makes a claim against CRW and/or its professional employees, at law or otherwise, for any alleged error, omission, or other act arising out of the performance of our professional services, and the CLIENT fails to prove such claim or prevail in and adversary proceeding, the CLIENT shall pay all costs incurred by CRW and/or its professional staff in defending itself against the claim. Neither party shall be responsible to the other for lost revenues, lost profits, cost of capital, claims of customers, or other special, indirect, consequential, or punitive damages. 12. DISPUTES All disputes, claims, and causes one party makes against the other, at law or otherwise, including third party or “pass-through” claims for indemnification and/or contribution, shall be initiated, determined, and resolved by arbitration in accordance with the Construction Industry Arbitration Rules of the American Arbitration Association, and judgment upon the award rendered by the Arbitrator(s) may be entered in any court having jurisdiction thereof. 21 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal CRW Engineering Group, LLC Page 5 of 5 Standard Terms and Conditions (11/6/16) In the event that one party makes a claim against the other, at law or otherwise, and then fails to prove such claim, then the prevailing party shall be entitled to all costs, including attorneys’ fees incurred in defending against the claim. 13. CLIENT LITIGATION If CRW is requested to produce documents, witnesses, or general assistance pursuant to a litigation, arbitration, or mediation in support of CLIENT litigation to which CRW is not an adverse party, CLIENT shall reimburse CRW for all direct expenses and time in accordance with CRW’s current rate schedule. 14. MISCELLANEOUS a) This Agreement supersedes all other agreements, oral or written, and contains the entire agreement of the parties. No cancellation, modification, amendment, deletion, addition, waiver, or other change in this Agreement shall have effect unless specifically set forth in writing signed by the party to be bound thereby. Titles in this Agreement are for convenience only. b) This Agreement shall be binding upon and inure to the benefit of the parties hereto and their respective successors and assigns provided that it may not be assigned by either party without consent of the other. It is expressly intended and agreed that no third party beneficiaries are created by this Agreement, and that the rights and remedies provided herein shall inure only to the benefit of the parties to this Agreement. c) No waiver of any right or remedy in respect of any occurrence on one occasion shall be deemed a waiver of such right or remedy in respect of such occurrence on any other occasion. d) All representations and obligations (including without limitation the obligation of CLIENT to indemnify CRW in Article 10 and the Limitation of Liability in Article 11) shall survive indefinitely the termination of the Agreement. e) Any provision, to the extent it is found to be unlawful or unenforceable, shall be stricken without affecting any other provision of the Agreement, so that the Agreement will be deemed to be a valid and binding agreement enforceable in accordance with its terms. 15. NOTICES All notices required or permitted to be given hereunder, shall be deemed to be properly given if delivered in writing by hand, facsimile machine, e-mail, or express courier addressed to CLIENT or CRW, as the case may be, at the addresses set forth below, with postage thereon fully prepaid if sent by mail or express courier. All notices, correspondence, deliverables, and invoices shall be submitted to CLIENT as indicated in the signed letter agreement unless otherwise indicated below: ___________________________________ ___________________________________ ___________________________________ Attn: _______________________________ All Notices and correspondence shall be submitted to CRW as indicated below: CRW Engineering Group, LLC 3940 Arctic Boulevard, Suite 300 Anchorage, Alaska 99503 Attn: D. Michael Rabe, PE, Member Manager 22 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Scope of Work for DeerStone Consulting Contribution to Mary’s Igloo and Kawerak Proposal to BIA TEDC Grant Program DeerStone Consulting will work with all project participants and stakeholders, led by Mary’s Igloo Traditional Council (MITC) and Kawerak, Inc., to assist in the following tasks and deliverables: • Conduct a Feasibility Study to determine an appropriate business structure and define development options for Unaatuq, LLC. to optimize and manage the locally available energy resources—including possibly geothermal, solar, biomass, hydropower, and wind—for a sustainable economy and business foundation for Pilgrim Hot Springs. • Perform an Economic Analysis of the income potential and requirements for an energy provider as well as a load analysis for various prospective businesses and enterprises based on development scenarios identified by project participants and stakeholders. • Facilitate a Strategic Planning Session with Unaatuq Board members to define scenarios and energy development options for Pilgrim Hot Springs and more broadly for the Kawerak/Norton Sound region. Assumptions 1. DeerStone Consulting team includes Brian Hirsch, Devany Plentovich, Tashina Duttle, Alan Mitchell, and Peter Crimp. 2. DeerStone will work closely with designated staff from MITC and Kawerak. Amanda Toerdal or her delegate will be primary Point of Contact for MITC and Kawerak and Brian Hirsch or his delegate will be primary Point of Contact for DeerStone. 3. The Economic Analysis will be a stand-alone activity but will be included as a major component of the broader Feasibility Study. The Economic Analysis will primarily be a techno-economic optimization of various energy options designed to meet different load profiles based on expected business activity at Pilgrim Hot Springs. The Feasibility Study will incorporate the Economic Analysis and Strategic Planning outputs into a much broader review of development pathways, approaches, institutional considerations, and overall strategies to support goal- and value-driven development of Pilgrim Hot Springs to serve as a role model for the region. 4. All Unaatuq partners will strive to have at least one representative available for the Strategic Planning Session, which will take place in Nome, Alaska, as travel restrictions allow, including a site tour of Pilgrim Hot Springs. 5. The Strategic Planning Session will be for 2 days in-person (if possible, or 1 day if virtual), and will require 1-2 person-days of preparation and 1-2 person-days of follow-up from DeerStone (depending on if in-person or virtual), including a write-up of the event, follow up next steps, and outreach support to participants before and after. 23 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal List of Bering Strait Tribes Chinik Eskimo Community (aka Golovin) King Island Native Community Native Village of Brevig Mission Native Village of Council Native Village of Diomede (aka Inalik) Native Village of Elim Native Village of Gambell Native Village of Koyuk Native Village of Mary’s Igloo Native Village of Saint Michael Native Village of Savoonga Native Village of Shaktoolik Native Village of Shishmaref Native Village of Teller Native Village of Unalakleet Native Village of Wales Native Village of White Mountain Nome Eskimo Community Stebbins Community Association Village of Solomon 24 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal 2020 KAWERAK, INC. Organizational Chart NATURAL RESOURCES DIVISION Kawerak President Kawerak Natural Resources Committee Executive Vice President Vice President Natural Resources Administrative Assistant Land Management Services Program Director Eskimo Walrus Commission Director Eskimo Walrus Commission Reindeer Herders Association Subsistence Resources Director Special Projects Assistant Land Management Specialist I Land Management Specialist II (2) Revised: 8/2020 Natural Resources Specialist Social Science Program Director Ice Seal Committee, Northern Norton Sound Advisory Committee, SPSRAC Local Assistant II (Nome) Social Science Research Assistant Natural Resources Advocate Probate Specialist II AYKSSI Steering Committee, NPFMC Ecosystem Committee, WASC Representative, WALCC Steering Committee Marine Mammal Commission, ICC Alaska Board and Executive Committee, AK Seagrant Advisory Member, Arctic Marine Mammal Commission, Arctic Waterway Safety Committee, IPCOMM Member FSB Tribal Consultation Committee * Kawerak Board Committee * NR programs serve as support staff * NR staff serve on these boards or committees Color Key: Society for the Anthropology of Consciousness Environmental Program Director Environmental Assistant Brownfields Coordinator Village IGAP Programs Probate Specialist I Local Assistants I (Villages) Bering Strait/Norton Sound Migratory Bird Council, Alaska Migratory Bird Co-management Council Emergency Preparedness Specialist Energy Development Specialist Marine Advocate Pilgrim Hot Springs General Manager & Caretaker25 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal 26 of 101Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration at Pilgrim Hot Springs 2010 to 2014: Final Report Prepared by the Alaska Center for Energy and Power at the University of Alaska Fairbanks 27 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal i Acknowledgments Activities described in this report were made possible with funding from a variety of federal, state, local, private and tribal sources including the U.S. Department of Energy under DE- EE0002846, “Validation of Innovative Exploration Techniques, Pilgrim Hot Springs, Alaska” and DE-EE0000263, “Southwest Alaska Regional Geothermal Energy Project, Pilgrim Hot Springs, Alaska,” The Alaska Energy Authority through RSA R1108 and R1215, the City of Nome, Bering Straits Native Corporation, White Mountain Native Corporation, Sitnasuak Native Corporation, Potelco, Inc., and the Norton Sound Economic Development Corporation. The geothermal exploration described in this report required a significant amount of planning and organization, and would not have been possible without the generous support from numerous individuals and organizations. The importance of local people and groups on the Seward Peninsula to this project’s success cannot be overstated. The employees of Bering Straits Native Corporation were generous with never-ending useful snippets of local knowledge as well as logistical support, in addition to the monetary support already described. Assistance from Robert Bensin, Kevin Bahnke, Larry Pederson, Matt Ganley, and Jerald Brown was indispensable and deserves special mention. The support of staff from the Norton Sound Economic Development Corporation and the City of Nome, especially John Handeland of the Nome Joint Utility Service and Mayor Denise Michels, were instrumental in overcoming logistical and funding challenges. Unaatuq, LLC and its board of directors have continued to have the vision required to keep the project moving forward. Unaatuq board member Roy Ashenfelter provided logistical support and boat transport up and down the Pilgrim River. Mary’s Igloo Native Corporation, whose land abuts the hot spring property was an important project partner and allowed land access for project activities. Mary’s Igloo Native Corporation tribal member Dora Mae Hughes and her family members provided important cultural background and shared stories about the regional history. Louis Green Sr. willingly shared his knowledge of the Pilgrim Hot Springs site and its history and provided logistical support. Chuck Fagerstrom freely shared his knowledge and was always willing to share site history and stories. Bryant Hammond and Amy Russell from Kawerak provided additional local support. University of Alaska faculty member Dr. Catherine Hanks assisted with technical editing and offered her expertise on the geology of the Seward Peninsula. Joe Batir and others from Southern Methodist University assisted with well logging and allowed ACEP to use high quality geothermal logging equipment. Jo Price and Graphite One Resources willingly shared data that they acquired to assist with the development of a regional geothermal understanding. In addition, former University of Alaska faculty members Dr. Ronald Daanen and Jo Mongraine were heavily involved and instrumental in project planning and data collection. The staff of the U.S. Geologic Survey (USGS) assisted in a variety of ways. Art Clark and the USGS drilling team made the initial stages of the project possible. John Glen and his crew accomplished an extensive set of geophysical surveys and interpretations and provided technical assistance at various stages of the project. The University of Alaska Fairbanks Geophysical Institute including Anupma Prakash, Christian Haselwimmer and Jeff Benowitz and graduate students Josh Miller and Arvind Chittambakkam 28 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal ii worked tirelessly to complete the remote sensing and conceptual modeling, adding an important piece to the body of knowledge about Pilgrim Hot Springs. Ryan Purcella of Baker Hughes and Mark Kumataka provided valuable engineering guidance related to well pumping and flow testing. Additional technical assistance was provided by Bill Cummings, and Dr. Dave Blackwell. Cheryl Thompson, collections assistant at the Carrie M. McLain Memorial Museum in Nome, was extremely helpful, providing assistance with historical research and obtaining historic photos. Ethan Berkowitz assisted with organizing and maintaining positive momentum during the final round of drilling and Howard Trott contributed his time and supplied equipment used in the flow testing. Joel Renner and Fran Pedersen spent long hours on a technical review of this report for which we are very grateful. A special thank you goes out to Dick Benoit who provided endless advice and technical assistance and who was always willing to answer his phone when we called. 29 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal iii Table of Contents Acknowledgments ....................................................................................................................... i List of Figures ............................................................................................................................. v List of Tables ............................................................................................................................. vi Terms and Acronyms ................................................................................................................ vii 1. EXECUTIVE SUMMARY .................................................................................................... 1 2. BACKGROUND – KRUZGAMEPA HOT SPRINGS .......................................................... 2 2.1 Geothermal Exploration History ....................................................................................... 4 3. GEOLOGIC SETTING .......................................................................................................... 6 3.1 Regional Geologic Setting ................................................................................................ 6 3.2 Local Geology ................................................................................................................... 8 4. SUBSURFACE TEMPERATURES ...................................................................................... 8 4.1 Updated Temperature Logging ....................................................................................... 11 5. REMOTE SENSING ............................................................................................................ 14 5.1 Satellite-based Geothermal Anomaly Mapping .............................................................. 15 5.2 Airborne Forward Looking Infrared Surveys ................................................................. 17 6. GEOPHYSICAL SURVEYS ............................................................................................... 27 6.1 Gravity Surveys .............................................................................................................. 28 6.2 Airborne Magnetic and Electromagnetic Surveys .......................................................... 29 6.3 Magnetotellurics Survey ................................................................................................. 31 7. DRILLING ACTIVITIES ..................................................................................................... 35 7.1 Permitting ........................................................................................................................ 36 7.2 Legacy Wellhead Repairs ............................................................................................... 36 7.3 Shallow Temperature Survey .......................................................................................... 37 7.4 Deep Drilling .................................................................................................................. 40 8. WATER CHEMISTRY ........................................................................................................ 42 9. FLOW AND INTERFERENCE TESTING ......................................................................... 44 9.1 Interference Testing of Wells PS-3, PS-4, and MI-1 ...................................................... 45 9.2 Interference Testing of PS-3, PS-13-1, and PS-13-3 ...................................................... 46 9.3 Flow Testing of PS-13-1 ................................................................................................. 46 9.4 Temperature and Pressure Monitoring in PS-13-2 ......................................................... 52 9.5 Temperature and Pressure Monitoring in PS-13-3 ......................................................... 52 9.6 Historic Hot Springs Temperature Monitoring ............................................................... 54 9.7 Flow Testing Conclusions .............................................................................................. 55 30 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal iv 10. PILGRIM GEOTHERMAL SYSTEM CONCEPTUAL MODEL .................................... 56 10.1 Conceptual Model History ............................................................................................ 56 10.2 Current Pilgrim Geothermal System Understanding .................................................... 61 11. EXPORTING GEOTHERMAL ENERGY TO NOME ..................................................... 64 11.1 Geothermal Power Economics ...................................................................................... 64 11.2 Wind-Diesel-Geothermal Microgrid Modeling ............................................................ 64 11.3 Transmission from Pilgrim Hot Springs to Nome ........................................................ 65 12. LESSONS LEARNED ....................................................................................................... 65 13. CONCLUSIONS ................................................................................................................ 66 14. REFERENCES ................................................................................................................... 68 APPENDICES APPENDIX A Well Schematics APPENDIX B Well Temperature Profiles APPENDIX C Well Locations and Descriptions APPENDIX D Wellhead Repair Description APPENDIX E Geophysical Survey Report APPENDIX F 2012 Mud Logging Records APPENDIX G Geophysical Well Logs for 2011 and 2012 Drilling APPENDIX H September 2013 Interference Testing APPENDIX I February 2014 Interference Testing APPENDIX J 2012 Drilling Logs APPENDIX K 2013 Geophysical Logs APPENDIX L Fugro MT Report APPENDIX M A Conceptual Model of Pilgrim Hot Springs: Joshua Miller Master Thesis APPENDIX N Reservoir Simulation Modeling: Arvind Chittambakkam Thesis APPENDIX O Tectono-thermal History of Pilgrim Hot Springs, Alaska APPENDIX P Wind-Geothermal-Diesel Microgrid Development: Jeremy VanderMeer Thesis APPENDIX Q Fuel Oil Volatility – Complications for Evaluating a Proposed Power Purchase Agreement for Renewable Energy in Nome, Alaska APPENDIX R High Voltage Direct Current Transmission Assessment at Pilgrim Hot Springs APPENDIX S Wind-Geothermal-Diesel Microgrid Development 31 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal v List of Figures Figure 1. The location of Pilgrim Hot Springs on the Seward Peninsula. ...................................... 2 Figure 2. Index maps showing the topography and regional geology. ........................................... 6 Figure 3. Topographic map of the area surrounding Pilgrim Hot Springs, .................................... 7 Figure 4. Map of all drill holes and well locations ....................................................................... 10 Figure 5. The1982 temperature logs from the original wells ........................................................ 11 Figure 6. Temperature profiles of all holes and wells .................................................................. 12 Figure 7. Map showing the approximate margin of the very shallow thermal aquifer ................. 13 Figure 8. Plan view temperature maps of Pilgrim Hot Springs .................................................... 14 Figure 9. A time series of ASTER visible to near-infrared imagery ............................................ 16 Figure 10. A subset of an ASTER wintertime false color composite image ................................ 16 Figure 11. Landsat 7 satellite images of the Pilgrim Hot Springs ................................................ 17 Figure 12. Low-emissivity thermal blankets ................................................................................ 18 Figure 13. Field calibration and validation data sites for the primary target area ........................ 19 Figure 14. Comparison of a FLIR-derived temperatures profile .................................................. 19 Figure 15. Mosaicked FLIR surface temperature data.................................................................. 20 Figure 16. FLIR (left) and optical data (right) from the fall 2010 survey .................................... 21 Figure 17. Processed airborne images for parts of the study area ................................................ 22 Figure 18. A simplified conceptual model of the Pilgrim geothermal system ............................. 23 Figure 19. A total surface energy budget model for the Pilgrim geothermal system ................... 24 Figure 20. The effect of wind speed on heat flux ......................................................................... 27 Figure 21. Gravity stations are labeled on a topographic map ..................................................... 28 Figure 22. Isostatic residual gravity map ...................................................................................... 29 Figure 23. Magnetic field maps from Glen et al. (2014) .............................................................. 30 Figure 24. Magnetic lineations interpreted from maximum horizontal gradients ........................ 30 Figure 25. Airborne EM resistivity slices ..................................................................................... 31 Figure 26. Magnetotellurics site locations. ................................................................................... 32 Figure 27. Resistivity at Profile D from a 1D MT inversion. ....................................................... 32 Figure 28. Resistivity maps at 25 m and 50 m from the blind 3D MT inversion. ........................ 33 Figure 29. Resistivity maps at 100 m, 150 m, 200 m, and 300 m ................................................ 34 Figure 30. Resistivity maps at 400 m, 500 m, 750 m, and 1000 m .............................................. 35 Figure 31. Areas of leaking, scale, and corrosion are shown on PS-4. ......................................... 37 Figure 32. The PS-4 completed replacement valve installation. .................................................. 37 Figure 33. Installing Geoprobe holes at Pilgrim Hot Springs. ...................................................... 38 Figure 34. Location of Geoprobe holes and their temperatures in Fahrenheit at 60 feet. ............ 39 Figure 35. The temperature logs from all Geoprobe holes ........................................................... 40 Figure 36. The mixing trend between sodium and chloride is shown for all samples .................. 43 Figure 37. Chloride content is shown along with well temperature ............................................. 44 Figure 38. PS-3 downhole pressure during interference testing. .................................................. 45 Figure 39. PS-3 temperature response during 2013 interference testing. ..................................... 46 Figure 40. Surface equipment used for the airlift of PS-13-1 ....................................................... 47 Figure 41. Downhole pressure and temperature record of PS-13-1 .............................................. 49 Figure 42. PS-13-1 downhole pressure and temperature .............................................................. 49 Figure 43. Downhole pressure and temperature at the end of the second airlift .......................... 50 Figure 44. Detailed flowing and static logs from PS-13-1 ........................................................... 51 Figure 45. PS-13-2 pressure and temperature response during PS-13-1 flow testing. ................. 53 32 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal vi Figure 46. PS-13-3 pressure and temperature response during PS-13-1 flow testing. ................. 53 Figure 47. The historic hot spring pool ......................................................................................... 54 Figure 48. Hot spring pool temperatures during the September 2014 flow testing ...................... 55 Figure 49. Conceptual model from Miller et al. (2013a). ............................................................. 58 Figure 50. Regional conceptual model cartoon from Glen et al. (2014). ..................................... 59 Figure 51. The current conceptual model of Pilgrim Hot Springs ................................................ 63 List of Tables Table 1. FLIR heat flux estimates. ................................................................................................ 26 Table 2. Permits and approvals ..................................................................................................... 36 Table 3. Pilgrim Hot Springs well chemistry in PPM .................................................................. 42 Table 4. Well productivity data .................................................................................................... 52 33 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal vii Terms and Acronyms ACEP Alaska Center for Energy and Power ASTER Advanced Spaceborne Thermal Emission and Reflection Radiometer DOE U.S. Department of Energy NETL National Energy Technology Lab EM Electromagnetic ETM+ Enhanced Thematic Mapper FLIR Forward looking infrared radiometry gpm Gallons per minute IGRF International Geomagnetic Reference Field Kauweraq The region of the central Seward Peninsula (also spelled Kawerak) MHG Maximum horizontal gradient MT Magnetotellurics MWe Megawatt electric MWth Megawatt thermal PGS Pilgrim Geothermal System PHS Pilgrim Hot Springs SMU Southern Methodist University TG Temperature gradient UAF University of Alaska Fairbanks Unaatuq Inupiaq word (also spelled Oonuktuak) meaning hot water/ hot spring. Also refers to the group of Native Alaskan and non-profit organizations that own Pilgrim Hot Springs. USGS United States Geological Survey VNIR Visible and near-infrared 34 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 1 1. EXECUTIVE SUMMARY This document is the final report for the Pilgrim Hot Springs (PHS) geothermal exploration project, funded by the U.S. Department of Energy (DOE), The Alaska Energy Authority, the City of Nome, Bering Straits Native Corporation, White Mountain Native Corporation, Sitnasuak Native Corporation, Potelco, Inc., and the Norton Sound Economic Development Corparation. The first round of funding in 2009 was awarded under Alaska Energy Authority RSA R1108 and R1215 and DOE award DE-EE0002846. In 2013, DOE award DE-EE0000263 along with match money from the six other organizations listed above was awarded. This report details the activities that occurred as part of the first and second rounds of funding for geothermal exploration at PHS in 2010 and 2013. The project objectives were to test innovative geothermal exploration techniques for low-to-moderate-temperature geothermal resources and conduct resource evaluations of PHS. A variety of methods including geophysical surveys, remote sensing techniques, and heat budget modeling estimated that the geothermal resource might support electrical power generation of approximately 2 MWe using a binary power plant. Further flow testing of the deep geothermal aquifer is needed to verify this estimate. Eight new wells were drilled around the PHS site to a maximum depth of 1294 feet. Five of these wells use sealed casing and can be used only to collect temperature logs. The other three wells have perforated casing and are capable of measuring temperature as well as artesian flow. A maximum temperature of 91°C (196°F) was measured in two different wells: in the shallow thermal aquifer at approximately 120 feet in depth and in the deep aquifer at approximately 1100 feet in depth. These wells were drilled in what is believed to be the vicinity of the upwelling zone, but both wells show a temperature reversal between the shallow and deep thermal aquifers, suggesting they are not directly over the main area of upwelling. Based on data collected to date, the main upwelling zone is likely northwest of well PS-13-1 in a swampy area that has been inaccessible for drilling. As in past surveys, geothermometry from water samples collected suggests maximum system temperatures could be as high as 145°C (293°F), based on Na-K-Ca geothermometry. The most concentrated geothermal fluid with 3500 ppm (parts per million) chloride continues to be collected from the traditional thermal hot spring located directly south of the church. Thermochronology data analyzed by University of Alaska Fairbanks (UAF) researchers suggest that the Pilgrim geothermal system (PGS) is relatively young, and core samples collected from the drilling indicate that temperatures have likely reached approximately 150°C (302°F) in the past 1000 years. In 2014, a power purchase agreement was signed between the City of Nome and Pilgrim Geothermal LLC, who has sent a letter of intent to the landowners to develop the resource. Modeling by the UAF power integration program, examined the effect of adding a geothermal generation source to the existing wind-diesel islanded grid in Nome. Adding 2 MW of geothermal power to the Nome grid displaces approximately 1 million gallons of diesel fuel per year (VanderMeer and Mueller-Stoffels, 2014). 35 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 2 Flow testing of the shallow thermal aquifer reached maximum flow rates of 350 gpm (gallons per minute), and sustained flow rates of 300 gpm for 7.5 hours. Based on the observed flow rates and minimum pressure decline, it appears likely that the shallow thermal aquifer could sustain this flow long term, opening up the potential for on-site direct geothermal heating or electrical power generation. 2. BACKGROUND – KRUZGAMEPA HOT SPRINGS Pilgrim Hot Springs, formerly known as Kruzgamepa Hot Springs, is located on Alaska’s Seward Peninsula about 60 miles north of Nome and 75 miles south of the Arctic Circle (Figure 1). The site has a long, colorful human history, which has included use as a traditional Native Alaskan gathering place, a farm, a dancehall and roadhouse, a Catholic orphanage and mission, and most recently as a recreational bathing and hunting site. The lush and tall local vegetation, dominated by cottonwood trees, contrasts with the otherwise treeless tundra of the western Seward Peninsula and is visible from miles away. Since the late 1970s, the area has seen two extensive geothermal exploration efforts that have extended road access to the site from the Nome-Taylor Highway. Before outsiders came to the region, the people of Kauweraq (the region of the central Seward Peninsula) used the area known as Oonuktuak (also spelled Unaatuq), also known as Kruzgamepa and later as Pilgrim Hot Springs. Traditionally, the hunting camp served as a Figure 1. The location of Pilgrim Hot Springs on the Seward Peninsula. 36 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 3 tropical oasis for the Kauweramuit (the people of Kauweraq) (D. Hughes, personal communication, February 17, 2015). During the winter months, Oonuktuak was an ideal living area with fresh water, plenty of wood for heat, and bountiful hunting and fishing. After successful ceremonial caribou hunts, other native groups such as the King Islanders would visit the area (Ray, 1992). Pressure from commercial whaling and hunting significantly reduced the marine mammal population in the region, and local government officials became concerned about the well-being of the region’s native inhabitants. In 1892, reindeer from northeastern Siberia were first introduced, after Dr. Sheldon Jackson, the Commissioner of Education in Alaska, received congressional approval. A reindeer station was established at Teller, about 40 miles west of the hot springs (Bucki, 2004). Later reindeer fairs were held, the first of which took place in 1915 at Pilgrim Hot Springs (Van Stone et al., 2000). Modern development at the hot springs began around the year 1900, during the Nome Gold Rush, when a family homesteaded 160 acres and worked the land, raising cows, chickens, pigs, and horses (Bland, 1972). After several years, the land was leased or sold to a series of people who developed a roadhouse. During the gold rush period, a bathhouse, greenhouse, roadhouse (hotel), and stables were built on the site. The facilities were frequented by the miners, their “fancy ladies,” and gamblers who reached the area by dog team. A railroad once passed within 8 miles of the site. In 1908, the roadhouse and saloon-dancehall burned to the ground. By this time, the gold rush was ebbing and a second roadhouse was constructed to serve travelers (National Register of Historic Places, 1977). By the late 1910s, mining on the Seward Peninsula had greatly diminished, and eventually, after another series of transactions, the land was deeded to the Catholic Church by two brothers with no heirs. In 1917 and 1918, an influenza epidemic decimated the area’s Native Alaskan adult population. On April 22, 1918, a Canadian priest and pastor of a Nome church, Father Bellarmine Lafortune, S.J, moved out to the hot springs to build an orphanage (The Alaskan Shepherd, 2009). Many buildings were moved from a mission that existed in the village of Mary’s Igloo, several miles north of the hot springs, to the present-day site. Additional buildings were constructed using lumber from a nearby mining site as well as the on-site timber (National Register of Historic Places, 1977). During this time, the site became known as Pilgrim Hot Springs (PHS). Eventually the orphanage included a machine shop, student dormitories, nun and priest quarters, a sizable church that now dominates the site, a variety of lesser buildings, a cemetery, and reportedly an unmarked or lost burial ground where victims of the Spanish influenza outbreak were interred. Some of the buildings were reportedly heated with the natural springs; others were heated using wood stoves. Historic photos show huge piles of firewood stacked near the church and a substantial treeless area around the springs, now heavily wooded. Toward the latter stages of the orphanage, firewood became scarce in the region (The Alaskan Shepherd, 2009). The orphanage was largely self-sustaining thanks to the gardens that flourished on the permafrost- free soil, producing legendary crops of potatoes, cabbages, turnips, and other vegetables. The population averaged about 100 youth and 20 adults then (National Register of Historic Places, 1977). A field of shoulder-high oats was growing in the thawed area in September 1915 (Waring, 1917). The orphanage closed in 1941; however, caretakers continued to grow produce, and up to 37 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 4 7 airplane flights a day ferried this produce to Nome (Bland, 1972), as there was no road access at the time. During World War II, military forces used the site for rest and recreational purposes. In 1969, Pilgrim Springs Ltd. signed a 99-year lease on the property with the Catholic Church to develop the site as a historical resort (Bland, 1972). This plan never materialized, but the land continued to be farmed by a number of caretakers. In 2010, Unaatuq LLC, a consortium of Alaska Native and nonprofit entities from the Seward Peninsula, purchased the property and decaying buildings for $1.9 million from the Fairbanks Catholic Diocese after the Diocese filed for Chapter 11 bankruptcy (Smetzer, 2010). Since acquiring the property, Unaatuq has been investigating various options for the development and preservation of the site. Throughout the site’s history, it has continuously been used for bathing and recreational purposes. 2.1 Geothermal Exploration History The first recorded description and map of the hot springs dates from 1915, after the local area had already seen significant development (Waring, 1917). Waring apparently reached the site via light carts pulled by dog teams on the old railroad grade that passed 8 miles east of the hot springs. Waring described a permafrost-free area 100 yards wide and a half-mile long, and measured a maximum spring temperature of 156°F. The visible single-point discharge in 1915 was only about 8 gpm, but additional diffuse discharge increased this amount to an estimated 60 gpm. The water was reported clear with a slight hydrogen sulfide odor. Waring collected a thermal water sample for chemical analysis. This analysis, now a century old, is remarkably similar to modern analyses of the thermal water (Table 3). In 1968, the Catholic Church leased the geothermal rights to C. J. Phillips of Nome (Kirkwood, 1979). However, no significant exploratory work occurred under this lease, which ultimately was revoked. The U.S. Geological Survey (USGS) designated the hot springs as a Known Geothermal Resource Area in the 1970s. In the early 1970s, initial evaluation of the geothermal resource commenced. The USGS revisited some of the thermal springs in central and western Alaska and published a new chemical analysis of the PHS thermal water (Miller et al., 1975). The quartz and Na-K-4/3Ca geothermometers from this analysis predicted subsurface hot springs temperatures of 137°C and 120°C. In October 1973, Harding-Lawson Associates ran two resistivity lines and concluded that a fault crossed the area and down-dropped bedrock from a depth of 100 feet to 600 feet (Kirkwood, 1979). In 1974, a 2250-foot-long north–south seismic refraction line and surface magnetic profile were run (Forbes et al., 1975). Forbes et al. measured a maximum temperature of 80°C in the thermal pool and deployed a portable seismograph for two nights to try to detect any tremors. They found the area quiet. The first major geothermal studies at PHS were led by the Geophysical Institute at the University of Alaska, the Alaska Division of Geological and Geophysical Survey, and the State Division of Energy and Power Development in 1979, using funding from the Alaska Division of Energy and Power and the U.S. Department of Energy. During the 45-day field season, a variety of geological, geochemical, geophysical, hydrological, and shallow drilling studies were performed at the site (Turner and Forbes, 1980). In the fall of 1979, the first two wells at PHS were drilled 38 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 5 to a maximum depth of 160 feet and initially flowed, allowing the first analysis of subsurface thermal water (Kline et al., 1980). During a 30-day field season in 1980, the central Seward Peninsula was evaluated on a more regional scale for its geothermal potential (Wescott and Turner, 1981). This helicopter-supported work included geologic, geophysical, and geochemical studies near PHS. It also incorporated a remote sensing component (Dean et al., 1982). In 1982, a 7-mile-long road was at last constructed from the Nome-Taylor Highway to Pilgrim Hot Springs, allowing reasonable access for a drilling rig and associated equipment capable of drilling larger-diameter and deeper wells (P. Eagan, personal communication, April 29, 2015). During summer 1982, four wells were drilled to a maximum depth of 1001 feet (Kunze and Lofgren, 1983; Lofgren, 1983). These wells were flowed, brief interference tests were conducted, and chemical analyses were obtained (Economides, 1982; Economides et al., 1982). This work represented the end of the first major exploration effort at PHS, as the maximum measured well temperature of 91°C was far too low for electrical power generation with the technology that existed at the time. Pilgrim Hot Springs attracted very little geothermal interest between 1983 and the early 2000s, with the exception of a comprehensive water and gas sampling program conducted in 1993 (Liss and Motyka, 1994). In the early 2000s, interest in the PGS gradually revived with the National Renewable Energy Laboratory sponsoring a site visit (Huttrer, 2002) and the Alaska Energy Authority funding a preliminary development feasibility study (Dilley, 2007). In 2008, the Nome Region Energy Assessment concluded that geothermal energy was a potentially economic option for the region (Sheets et al., 2008). In 2006, the first geothermal power plant in Alaska was installed at Chena Hot Springs, near Fairbanks. The project was able to generate electricity using 165°F (73°C) fluid, effectively making it the lowest temperature geothermal power plant in the world and demonstrating that generating electricity from low temperature geothermal resources was technically and economically feasible (Holdmann, 2007). Following this success, overall interest in developing Alaska’s low-to-moderate temperature resources increased, and the Alaska Center for Energy and Power (ACEP) secured grant funding from the U.S. Department of Energy and the Alaska Energy Authority to resume exploration of the PGS. This work began in 2010, with repairs to the existing wellheads so that those wells could be relogged and flow tested. Remote sensing studies also began at this time (Haselwimmer et al., 2011), followed by numerical modeling of existing data (Daanen et al., 2012). The USGS also collected additional geophysical data around the hot springs (Glen et al., 2012). In 2011, two 500-foot temperature gradient holes were drilled to evaluate the northern part of the thermal anomaly where the thermal upwelling was then expected to be located. In 2012, three deep holes were drilled in an attempt to precisely define the location of the thermal upwelling beneath the shallow thermal anomaly (Miller et al., 2013a; Miller et al., 2013b; Benoit et al., 2014a). Recent modeling efforts used data from the deep holes drilled in 2012 (Chittambakkam et al., 2013). In 2013, additional funding became available through the U.S. Department of Energy. ACEP drilled a deep, large-diameter well and two shallower wells 39 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 6 to locate and produce fluid directly from the deep thermal upwelling (Benoit et al., 2014b). These holes failed to penetrate or precisely locate the thermal upwelling, but were completed as possible future production wells for direct-use projects. In September 2014, these three wells were flow tested and monitored for interference. 3. GEOLOGIC SETTING 3.1 Regional Geologic Setting The central Seward Peninsula is underlain by a Precambrian metamorphic complex, intruded by Cretaceous granitic rocks (Amato and Miller, 2004; Till et al., 2011). In the vicinity of Pilgrim Hot Springs (PHS), Quaternary alluvial fill overlies this basement complex. The metamorphic and intrusive rocks are well exposed in the Kigluaik Mountains 2.5 miles south of the PGS and on Mary’s Mountain and Hen and Chickens Mountain 2.5 miles north of the hot springs (Figure 2). Nowhere is the alluvial fill dissected to the point that any meaningful thickness can be viewed in any detail at the surface. The dominant regional structural feature near PHS is the east–west trending Kigluaik- Bendeleben system of normal faults. These normal faults are interpreted as due to regional north- south extension (Ruppert, 2008), which led Wescott and Turner (1981) to propose that the central part of the Seward Peninsula is a 250 km long east-west striking rift system. The Kigluaik section of the fault system uplifts the Kigluaik Mountains in the south relative to the Imuruk Basin in the north, where the hot springs are located. Hudson and Plafker (1978) divided the Kigluaik section of the fault system into three segments. The western and central section’s show clear surface traces and post Wisconsin or Holocene vertical displacements up to 10 m. The eastern section, which passes about 2.5 miles south of PHS, has less definable surface traces, being more obscured by glacial deposits (Hudson and Plafker, 1978). The eastern section is more complex, with two distinct northward steps, giving Figure 2. Index maps showing the topography and regional geology.The red box in the left panel shows the area in the geologic map on right. The location of Pilgrim Hot Springs is shown by the red star. Map after Till et al. (2011). The red box in the right panel outlines the area of Figure 3. 40 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 7 the range front an en echelon system of at least three mappable faults. Geomorphic features suggest that displacement of the western en echelon section is younger than displacement of the eastern section. The vertical displacement on the Kigluaik fault zone is at least several hundred meters and probably exceeds 1200 m. While it is tempting to hypothesize that this major extensional structure somehow plays a role in the geothermal system, no serious arguments for this have yet been made. Up to 320 m of Quaternary alluvium ranging from clay to gravel in size and consisting of alluvial, fluvial, glaciolacustrine, and brackish lagoon sediments has been drilled in the immediate vicinity of PHS (Miller et al., 2013a, 2013b). The volcanic rocks closest to the PGS are the Holocene Lost Jim basaltic lava flows (Till et al., 2011). These flows cover 88 square miles and lie about 30 miles northeast of the PGS. They are outside of the boundary of Figure 2. This distance from the PGS makes it unlikely that the Lost Jim lava flows represent a possible direct magmatic heat source for the PGS. The northern horn of the Seward Peninsula also hosts the world’s largest maar craters, dated at 21,000 years (Rozell, 2006), but these craters are much farther away. If there is a magmatic heat source for the PGS, no author has yet tried to make a convincing case for its existence Figure 3. Topographic map of the area surrounding Pilgrim Hot Springs, indicated by the red star. 41 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 8 3.2 Local Geology The localized surficial geology surrounding the PGS consists of a flat, wide valley covered mostly by thermokarst lakes, permafrost tundra, and muskeg swamps (Miller et al., 2013b). The most striking local surface feature is a thaw in the permafrost, covering an area of about one-half square mile (~0.58 mi2 or 1.5 km2) that allows anomalous vegetation such as cottonwood trees, alders, grass, and various wildflowers to thrive. The most extensive published surficial geology description of the PGS suggests that the hot springs might be located near the western edge of an actively subsiding north-south striking graben, apparently resulting from north–south-trending faults (Kline et al., 1980). The publication offers eight brief lines of evidence as support, but unfortunately, contains no maps, photos, and/or diagrams to back the evidence, nor has any been reported in more recent publications. Swanson et al. (1980, p.11) suggest that “many of the canyons found on the north flank of the Kigluaik Mountains are apparently controlled by north–south-trending faults.” However, the 2011 geologic map of the Seward Peninsula (Till et al., 2011) shows no north– south-trending faults in the Kigluaik Mountains, casting serious doubt on the earlier suggestion. In spite of the 2011 map, inferred or buried north–south-trending faults are shown by Miller et al. (2013a) and are included in discussion by Glen et al. (2014). Thus, at this time, north–south- trending structures have been proposed by several researchers, but no recent geological work has focused on the Pilgrim Valley to confirm the existence of these structures, and a more recent geologic map did not give them any credence. On a smaller scale, the local geology has been evaluated with several recent drill holes to a maximum depth of 350 m (Miller et al., 2013a, 2013b). This evaluation primarily focused on the stratigraphy of the Quaternary alluvium and showed that metamorphic bedrock is present at a depth of about 320 m. Particle sizes in the alluvium range from clay to gravel, with sand, silt, and clay predominating. The sand is locally indurated with silica cement near most of the deeper wells that have been drilled. The most laterally extensive silt and clay unit is located about 164 feet (50 m) above the top of the metamorphic basement. 4. SUBSURFACE TEMPERATURES Above the top of the metamorphic bedrock at depths of about 1050 feet (320 m) the thermal fluid flow pattern has become much better defined by the activities described in this report. Some type of vertical or near-vertical permeable channel allows the thermal fluid to rise to the surface through a sequence of unconsolidated Quaternary fluvial material. If any elongation or dip accompanies this channel, it has not yet been recognized. It is possible that the access limitations for drill-hole locations allow some northwest–southeast elongation, which could be hypothesized as evidence for a fault. Drilling and temperature logging completed between 1979 and 2014 have delineated a 2 square mile permafrost-free area and a series of thermal aquifers overlying each other within this location. All holes and wells that were drilled between 1979 and 2013 are shown in Figure 4. The oldest well logs are from September 1982, when flowing and static temperature logs were obtained from the first six wells using a FENWAL model UUT-51J1 thermistor instrument, with 42 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 9 an estimated absolute accuracy of about 1°C (Lofgren, 1983) (Figure 5).A few logs from this era are questionable, especially below a depth of 200 feet in the PS-5 well log, but overall these well logs give a valuable baseline dataset with which to measure long-term aquifer temperatures. All existing and new holes and wells, except for PS-2, were repeatedly logged between 2011 and 2014. In general, temperature profiles matched the profiles reported in Woodward-Clyde (1983). Recent well logs were obtained using two different instruments. Many logs were obtained using one of three Kuster K-10 memory tools owned by the Alaska Center for Energy and Power. The Kuster tools are extremely robust, can be used up to 150°C (302°F) and 5000 psi, and can remain downhole for long periods. This tool measures temperature and pressure with an accuracy of 0.2°C. The Kuster tools were used with a strong reel of aircraft cable that could be operated by hand by one person. Once retrieved from the hole, the Kuster tool is disassembled and the data are downloaded onto a computer. 43 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 10 Figure 4. Map of all drill holes and well locations at Pilgrim Hot Springs. 44 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 11 The second tool was a light and simple portable wireline temperature measurement tool with surface readout, custom built for Southern Methodist University (SMU). The tool employs a platinum thermistor with a reported accuracy of ±.01°C depending on the depth. It is simply referred to here as the SMU tool. Both types of equipment were compared with one another, and the readings were virtually identical. 4.1 Updated Temperature Logging A presentation of all available PHS temperature profiles to a depth of 160 feet (Figure 6) reveals a confusing picture, but highlights a hot shallow aquifer of varying depths. The thermal anomaly consists of a shallow aquifer 10 to 20 feet deep (Figure 7) above an aquifer 55 to 90 feet deep, which is referred to here as the shallow thermal aquifer (Benoit et al., 2014a). The shallow thermal aquifer is the primary geothermal discharge zone of the geothermal fluid within the PGS. The shallow thermal aquifer can be subdivided into northern and southern portions based on the shape of the static temperature profiles measured in the associated holes and wells. While these northern and southern shallow thermal aquifers have different characteristics, they are likely not independent aquifers and are certainly hydraulically connected. Where the holes and wells penetrate the shallow thermal anomaly and show a temperature reversal, the temperature profiles define the aquifer temperature, depth, and thickness. These data were used to create an aquifer temperature map showing the flow direction and the division of the northern and southern shallow aquifers (Figure 8). Figure 5. The1982 temperature logs from the original wells drilled at Pilgrim Hot Springs from 1979–81. 0 100 200 300 400 500 600 700 800 900 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 Depth (feet) Temperature (F) Pilgrim Hot Springs Static Temperature Logs PS-1 PS-2 PS-3 PS-4 PS-5 MI-1 45 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 12 Figure 6. Temperature profiles of all holes and wells drilled to date at Pilgrim Hot Springs are shown. The temperature profiles in the top graph define the shallow and very shallow thermal aquifers shown in Figure 8 and Figure 7. The bottom graph shows all the deep holes and wells drilled to date. These profiles show the temperature minimum data that were used to create Figure 8. 0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 20 40 60 80 100 120 140 160 180 200 Depth (feet) Temperature (F) All Pilgrim Hot Springs Static Temperature Logs 0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 40 60 80 100 120 140 160 180 200 Depth (feet) Temperature (F) Pilgrim Hot Springs Deeper Static Temperature Logs S1 S9 PS-3 1982 PS-4 1982 PS-5 1982 MINC-1 1982 PS 12-1 2013 PS 12-2 2013 PS 12-3 2013 PS 13-1 Combined PS 13-2 Combined PS 13-3 10-29-13 46 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 13 The hottest measured temperatures in the shallow aquifer occur near the boundary of the northern and southern thermal aquifers, indicating that both are supplied from the same upwelling source and represent thermal fluid moving through permeable intervals of varying thicknesses and depths. Overall, temperature distribution of the shallow thermal aquifer appears to show primitive waters rising in its center and flowing out laterally (Figure 8). The weaker thermal aquifers penetrated by the remote northeasterly S1 and S9 holes are most likely a continuation of the northern aquifer The deep holes that have been drilled at PHS show temperature minimums in between depths of 220 and 400 feet (Figure 6). While only the 12 deepest holes penetrate the temperature minimum to a depth where positive gradients occur, they allow the creation of plan view map showing temperature minimum contours (Figure 8). The temperature minimums measured between the shallow and deep thermal aquifers, and shown in Figure 6, provide the best dataset to define the location of the upwelling zone in Figure 8. Since 91°C (196°F) fluid has been measured in the deep thermal aquifer at the top of bedrock, and 91°C fluid has been measured in the shallow thermal aquifer, there must be a zone where the 91°C fluid emerges from a fracture of some type in the metamorphic bedrock and travels between those two aquifers. The two maps in Figure 8 suggest that the thermal fluid is rising through bedrock in the northwest swampy area and flowing northeast and south through the shallow thermal aquifer. The temperature profile from well PS-12-2 shows identical temperatures of 90°C (194°F) at depths of 126 feet and 1148 feet, indicating that the geothermal fluid loses no heat as it rises from the top of bedrock to the shallow thermal aquifer. Therefore, we speculate that the hottest and most saline fluid samples collected from the thermal springs and the shallow thermal aquifer have probably not been diluted by any shallow groundwater within the unconsolidated Figure 7. Map showing the approximate margin of the very shallow thermal aquifer, the temperatures within this aquifer, and the temperatures of thermal water measured at the surface. The red boundary closely approximates a temperature of 80°F (27°C). 47 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 14 Quaternary fluvial material. This indicates that pressures are higher within the plumbing hosting the thermal flow than in the surrounding Quaternary material. 5. REMOTE SENSING The first calculations of heat loss and potential power output of the Pilgrim geothermal system (PGS) were developed from 1979 data (Harrison and Hawkins, 1980; Osterkamp et al., 1980). Harrison and Hawkins (1980) indirectly measured the surface discharge downstream from the main area of PHS at 67 gpm and used a hot water temperature of 81°C to calculate admittedly crude numbers of 1.5 and 2.2 MW due to thermal water surface discharge. A 10 MW total vertical heat flow from the thawed area around the springs was also determined. Harrison and Hawkins (1980) indicate that this total vertical heat flow is probably a serious underestimate, as it did not include the power removed by groundwater movement. It is now known that 91°C would be a more accurate original thermal water temperature. Osterkamp et al. (1980), using a Figure 8. Plan view temperature maps of Pilgrim Hot Springs. The temperature contour map on the left shows the shallow thermal aquifer at the hot springs, and is based on the shallow temperature maximum. On the right, temperature minimum contours are shown at the hot springs. Known temperature contours are shown as solid lines. Hypothetical higher temperature contours are shown as the closely spaced dashed lines These minimum temperature contours, based on deep holes and wells, in conjunction with the shallow aquifer temperature profiles, indicate the direction of thermal water flow and help pinpoint the likely upwelling zone northwest of the area where past drilling occurred. 48 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 15 more systemwide approach to estimate the heat loss, analyzed the temperature and salinity increases in the Pilgrim River after it had passed through the geothermal area. This approach resulted in minimum total accessible power values of 350 to perhaps 500 MW. However, Osterkamp et al. admit that these numbers are highly uncertain, and caution that the values “should not result in unbridled optimism.” The first remote sensing efforts at PHS occurred in 1980 with radar and infrared surveys (Dean et al., 1982). The radar study identified numerous lineaments near the PGS that have received little or no recent attention. The high altitude (60,000 ft) infrared work indicated the presence of two large and unusually warm areas along the Pilgrim River north of the hot springs, but provided no quantitative thermal data. Remote sensing work since 2010 has extended the traditional use of remote sensing for geothermal exploration by developing methods for acquiring and processing remote sensing images (Haselwimmer et al., 2011). These methods identified various surface signatures associated with the geothermal systems and derived first-order quantitative estimates of thermal fluxes. Permafrost-free areas, snowmelt areas in early spring, anomalous vegetation patterns, and heated ground and water bodies were identified as areas that warrant further study. The temperature images derived from remote sensing provided the basis for heat budget modeling. This helped to focus the field efforts for further investigation and helped to target drilling activities and develop a conceptual heat flow model. 5.1 Satellite-based Geothermal Anomaly Mapping Satellite images from Landsat, Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER), and WorldView-2 (WV-2) were processed and used to identify areas of persistent high temperature, areas of snowmelt in winter images, and areas of greener vegetation in springtime images. An iterative approach to the use of satellite data followed by airborne surveys and traditional ground-based exploration was recommended as a routine part of a systematic geothermal exploration program. 5.1.1 Analysis of Landsat 7 Data A search of the Landsat 7 archive for ETM+ images from the PGS region yielded 18 scenes, which had been acquired between August 1999 and July 2010. Eleven datasets were selected for further analysis of cloud and snow-free images. The discrimination of thermal anomalies was undertaken using the image “stacking” approach (Prakash et al., 2011). This included pre-processing the band 6L thermal data for each dataset using the three-step procedure described by Chander et al. (2009). Thermal hot spot images for each year were integrated to identify temporally persistent thermal anomalies most likely to represent geothermal sources. A thermally anomalous pixel identified in data from three different years was labeled persistent. The ETM+ data highlight five persistent thermal anomalies located within the broad region of the PGS. These anomalies were later investigated in detail during the aerial FLIR survey. 49 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 16 5.1.2 Analysis of ASTER Data The ASTER multispectral thermal infrared data were acquired over the PHS site to identify geothermal anomalies (Figure 9). The 90 m spatial resolution of the ASTER thermal bands is lower than that of Landsat 7; however, as a multispectral instrument, ASTER is routinely used to acquire data during its nighttime ascending orbit, minimizing the effects of solar heating. The five ASTER thermal bands also enable the effects of emissivity to be accounted for within geothermal anomaly detection. The ASTER data delineated potential surface indicators of geothermal activity such as snowmelt anomalies, anomalous river ice melt, and areas of vegetation growth in the PGS region. Figure 9. A time series of ASTER visible to near-infrared imagery (top) and thermal (bottom) data from Pilgrim Hot Springs, showing snow-free areas and vegetation growth anomalies associated with geothermally heated ground. Figure 10. A subset of an ASTER wintertime false color composite image with 15 m spatial resolution is shown on the left. Prominent snow-free areas are indicated with red arrows. The left arrow points to the area near the hot springs. The right arrow points to a persistent snow- free region. A WV-2 color infrared image acquired in May 2010 is shown on the right. Healthy green vegetation (bright pink/reddish tones) and senescent vegetation (dark brownish red tones) are clearly visible (left). The processed WV-2 image (right) shows vegetation vigor, the dashed white line marking the approximate limit of vigorous nontundra vegetation. This map (right) is a color-coded Normalized Difference Vegetation Index (NDVI) image, where NDVI = (Near- infrared – Red) / (Near-infrared + Red). 50 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 17 5.1.3 Analysis of WorldView-2 Data Analysis of high-resolution VNIR (visible and near-infrared) data was completed using the commercial WV-2 satellite data. Images from WV-2 are acquired in the visible and near-infrared region of the electromagnetic spectrum at a spatial resolution of 1.2 m. The presence of the near- infrared band and high spatial resolution makes the dataset suitable for detailed vegetation mapping. Data were acquired during May 2010 and defined vegetation growth anomalies associated with geothermally heated ground (Figure 10). This work was validated with shallow- temperature survey measurements during the 2011 and 2012 field seasons that outlined the extent of the shallow and very shallow thermal aquifers (Benoit et al., 2014a). 5.2 Airborne Forward Looking Infrared Surveys Forward Looking Infrared Radiometer (FLIR) data collected from airborne surveys were used to calculate the geothermal potential of the PGS using a thermal budget model. Airborne surveys were conducted in fall 2010 and spring 2011, and data were mosaicked and processed to create high-resolution optical and thermal images. Thermal data-processing algorithms used by the volcanology community were adapted to compute heat flux. Airborne surveys were planned around high- and low-priority survey areas (Figure 11) to provide flexibility in case of poor weather conditions. The primary survey area covered a region approximately 27 km2, centered on the main PGS site encompassing the most likely geothermal anomalies detected from the Landsat 7 ETM+ data (red polygons in Figure 11). The secondary survey area covered a region approximately 175 km2, including the sites of the other thermal anomalies detected from Landsat. Figure 11. Landsat 7 satellite images of the Pilgrim Hot Springs region. The left image shows the extent of the primary and secondary survey areas; thermal anomalies detected from Landsat 7 satellite data are indicated by red filled polygons. On the right are the flight line locations for the aerial survey over the hot springs; cloud and turbulence restricted data acquisition over the southern half of the secondary survey area. 51 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 18 The first airborne survey was undertaken from September 9–15, 2010, using the Nome Airport as the base for flight operations. Favorable weather conditions enabled acquisition of data over the entire primary survey area and the northern portion of the secondary survey area. Flights over the southern portion of the secondary area were not possible due to persistent cloud cover and turbulence around the northern flanks of the Kigluaik Mountains. The FLIR images were successfully acquired along all the flight lines shown in Figure 11. Optical imagery was acquired for most of the flight lines; however, technical issues led to some gaps in the imagery in the northern part of the secondary survey area. Thermal images were acquired using a FLIR Systems A320 camera that records emitted thermal infrared radiation in the 7.5 to 13 µm wavelength region. The FLIR has a 320  240 pixel sensor with a 25 µm sensor pitch and 18 mm lens. Visible images were acquired using a Nikon D700 digital camera with an 85 mm f/1.8 lens fixed at infinity. The FLIR and D700 cameras were positioned side-by-side in a fixed nadir-looking mount within the aircraft. The FLIR camera was set to continuously record thermal images at a frame rate of 5 Hz, and Topoflight Navigator software triggered the shutter of the D700 camera at pre-programmed intervals along the flight lines. A Crossbow NAV440 GPS/IMU unit recorded the position, roll, pitch, and yaw of the plane during the survey. A flying height of about 1000 m yielded an approximate spatial resolution of 1.4 m for the thermal imagery and 20 cm for the optical imagery. The second airborne survey was flown in April 2011 and was restricted to a small area centered on the PGS property. During this survey, optical images were acquired at 20 cm resolution and FLIR data were acquired at 1.2 m spatial resolution. In- flight GPS data were recorded and time synced with the optical and FLIR image frames. 5.2.1 Field Calibration and Validation Concurrent with the fall 2010 airborne survey, a field party of three undertook ground calibration and validation work in support of the airborne FLIR and optical data collection. Accurate geographic positions of well-spaced and notable ground features and thermal blankets (Figure 13) were recorded using portable Garmin and Trimble GPS receivers. These ground control points enabled georegistration of the FLIR and optical data. Thermal blankets provided geo-located “cool” targets readily delineated from the FLIR data (Figure 12). Figure 12. Low-emissivity thermal blankets (cold targets) were used as ground control points for registration of airborne FLIR and optical image data. 52 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 19 Figure 13. Field calibration and validation data sites for the primary target area of the Pilgrim Hot Springs survey; the data are overlain on a high resolution color near-infrared aerial photograph (AHAP) of the study area. Wind speed, temperature, and humidity measurements were recorded throughout the collection period to calibrate the thermal data. Ground and water temperatures were recorded using TEGAM thermocouple sensors. Several ground temperature profiles were also recorded near the main hot spring site to compare against the retrieved FLIR surface temperature data, enabling further calibration as needed (Figure 13). Two HOBO temperature-logging systems provided continuous measurements of ground temperatures after the survey had been completed. The region around the main hot springs site and an area about 3.5 km northeast along the Pilgrim River, where field observations provided some evidence for a geothermal anomaly, were the priority regions (Figure 15). Initial processing of the FLIR data required knowledge of the surface temperatures and humidity values as inputs to the ThermaCam research software. The average flying height was also integrated to correct for atmospheric absorption and emission. A comparison of collected FLIR surface temperature values with ground-based temperature profiles shows agreement to within about 5°C (Figure 14). For the first airborne survey, the surface temperature images were manually georegistered to a high-resolution aerial photograph of the region from the Alaska High-Altitude Aerial Photography (AHAP) program and then mosaicked using ArcGIS software. There was significant overlap of the individual FLIR frames, associated with the 5 Hz acquisition rate. A high-quality mosaic Figure 14. Comparison of a FLIR-derived temperatures profile (black line) with a field temperature profile (red line) for a selected profile line. 53 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 20 image was obtained using every fifth image. Color in the visible images was adjusted to improve the contrast, then georeferenced against the AHAP aerial photograph and mosaicked together with a minimum of overlap. A semi-automated methodology was used to mosaic the spring 2011 images. The in-flight time synced GPS information was synchronized with the optical and FLIR sensor systems to georeference each image. To mosaic the images together, 2d3 software was used. Due to logistical challenges, the second round of field validation work was delayed until August 2011. This fieldwork included:  Gathering in situ measurements of hot spring temperatures.  Validating the locations of springs mapped from FLIR data, and acquiring in situ thermal images of hot springs and pools.  Measuring the outflow rate of hot springs.  Validating the extent of snowmelt anomalies and inferred geothermally heated ground using 1.20 cm shallow temperature probes.  Recording the temperature and conductivity of the Pilgrim River as well as local streams and locating outflow of saline geothermal waters. 5.2.2 Mapping Using Airborne Images The main surface geothermal features such as hot springs, wells, pools, and areas of hot ground can be clearly delineated using the fall 2010 FLIR imagery with its 1.3 m resolution. The surface water temperatures in the images are as high as 40.5°C (105°F). The FLIR imagery helped Figure 15. Mosaicked FLIR surface temperature data for the main Pilgrim Hot Springs site (bottom left) and possible geothermal area to the northeast (top right). 54 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 21 researchers to delineate geothermal features that would be difficult to map using visual imagery alone. Examples of such features include:  Upwelling thermal plumes within pools of water  Temperature gradients within pools and streams indicating the flow paths and mixing of hot and cool waters  Subtle thermal features that may represent previously unmapped small springs  Areas of hot ground away from the main spring complex. The optical images acquired during the same period provide useful complementary information, especially about land cover in the area (Figure 16). The analysis of the FLIR data from the area northeast of the main PHS site (Figure 15) provided little evidence for current geothermal activity. The range of surface temperatures is consistent with the different surface types (vegetation, soil, water ponds), and there are no obvious thermal anomalies. Nevertheless, the ground cover present in this region is similar to the ground cover near the hot springs, and it is not found elsewhere in the region. The April 2011 survey was completed in early spring when the region usually is still covered by a thick blanket of snow. The survey timing proved useful for mapping areas of snowmelt (Figure 17), a direct indicator of surface heating from the very shallow geothermal aquifer. Snowmelt areas also correspond to permafrost-free areas and anomalous vegetation growth not regularly found on the Seward Peninsula. The spring FLIR data were more useful than the fall FLIR data in identifying the limits of the very shallow thermal aquifer (Figure 17). Figure 16. FLIR (left) and optical data (right) from the fall 2010 survey over the main Pilgrim Hot Springs site. The FLIR data effectively delineate surface features associated with the geothermal system, such as hot springs, pools, and warm ground. 55 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 22 Figure 17. Processed airborne images for parts of the study area. Top left: Temperature map from September 2010 FLIR survey. Top right: Temperature map from April 2011 FLIR survey. The April 2011 image more clearly reveals the limits of the shallow hot aquifer. Bottom left: Subset of the April 2011 image, indicated with a white box in top right panel. Bottom right: Optical image of the area corresponding to the image in the bottom left panel. The optical image reveals underlying soils (brown), as the snow has melted over these areas due to geothermal heating. 56 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 23 5.2.3 Heat Budget Modeling A heat budget model was developed to quantify the radiant and convective heat flux and the flow rate of surface geothermal waters (Figure 18). An initial model treated all hot pixels in the same way, regardless of whether they were associated with heated ground or hot water. Upon further examination, it became clear that hot ground and hot water gain and lose heat differently, and the thermal flux estimations for these features need different approaches. An improved heat flux modeling process was developed (Haselwimmer et al., 2011; Haselwimmer and Prakash, 2011). Both approaches are discussed in this section. Initially, heat loss was estimated from the geothermal system by correcting for background temperature and the natural radiative heat loss of the earth and sun. Using a modified Stefan- Boltzmann equation (see below) with fixed values for surface emissivity and background temperature, the radiant flux was calculated for each pixel representing a geothermal feature: Figure 18. A simplified conceptual model of the Pilgrim geothermal system used for numerical calculations of thermal flux from the processed FLIR data. 57 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 24 Μ (Τh4 - Τb4) where To delineate the pixels associated with geothermal areas, a mask was created using a temperature threshold applied to the FLIR image. The background temperature value used in the thermal flux calculation was the average temperature value from the non-geothermal areas (not including anthropogenic and other non-geothermal temperature anomalies). The radiant flux value for each geothermal pixel was summed to calculate the total radiant flux, which amounted to 6.2  105 Watts. This method underestimated the thermal flux associated with the hot waters, so a sensitivity analysis was not performed (Haselwimmer et al., 2011). Upon further consideration, we concluded that the convective component was likely the dominant heat transfer component. Later model development attempted to establish methods for estimating the convective heat flux from geothermal hot springs and pools. Pixels associated with hot waters and hot ground are easily separated on the FLIR image mosaics. These water pixels were isolated for further analysis. Adapting an approach applied to volcanic crater lakes (e.g., Patrick et al., 2004), an energy budget model was developed to quantify the convective heat flux along with the flow rate of the surface geothermal waters at PHS (Figure 19). Complete details about the thermal model used for the quantitative analysis are presented in Haselwimmer and Prakash (2011) and are briefly Μ = radiant flux density (W/m2) ε = emissivity σ = Stefan-Boltzmann constant Τh4 = temperature of pixel in Kelvin Τb4 =temperature of background in Kelvin Figure 19. A total surface energy budget model for the Pilgrim geothermal system. Refer to the main text for an explanation of each term. 58 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 25 described below. The total heat budget for a water body (in Watts) expressed as Фtotal = Фgeo + Фppt + Фseep + Фevap + Фsens + Фrad + Фsun + Фsky where Simplifying this model further, Фppt and Фseep were removed, as these heat fluxes are small. The temperature of surface non-geothermal waters was used to account for Фsun and Фsky terms. Pixels associated with geothermal surface waters were isolated, and the geothermal heat flux density was calculated in W/m2on a pixel-by-pixel basis using the following equation: qgeo = (qrad + qevap + qsens) - (qradAmb + qevapAmb + qsensAmb) where qrad, qevap, qsens and qradAmb, qevapAmb, qsensAmb are radiative, evaporative, and sensible heat fluxes for each pixel at the ambient temperature of non-geothermal waters. Further, qrad, the radiative heat flux, was calculated using the Stefan-Boltzmann equation: qrad = εσT4 where Also, qevap+sens, the evaporative and sensible heat fluxes, were calculated using the formula presented by Ryan et al. (1974): qevap+sens = [λ(Tsv-Tav)1/3+ boW2][es-e2+C(Ts-Ta)] Фgeo = heat input from geothermal fluids Фppt = heat input from precipitation Фseep= heat flux from seepage Фevap= heat loss from evaporation Фsens= heat loss via sensible heat transfer Фrad = heat loss by radiation Фsun = heat input from solar radiation Фsky = heat input from atmospheric radiation σ = 5.67 x 10-8 (Stefan-Boltzmann constant in W/m2 K-4) ε = water emissivity(0.98) T = water temperature (°C). 59 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 26 where This model was applied to both FLIR datasets. The total heat flux is the sum of heat fluxes for each pixel, representing the hot water at the surface. Flow Rates Assuming a fixed hot springs temperature of 81°C and water at the ambient air temperature, the flow rate (V) in m3/s was calculated from the total geothermal heat flux (Фgeo) using the following equation: V = [Фgeo / (hs-hamb)] / ρw where Heat Budget Modeling Results The computed heat flux/flow rate estimates are generally higher than the in situ observations. This difference is likely caused by underestimating in situ measurements of the total outflow rate of the hot springs. These calculated results are quite conservative as they assume a wind speed of 0 m/s, which is unrealistic for the PHS area. The nearest meteorological station about 50 km northeast of PHS reports an average annual wind speed of 3.18 m/s. Therefore, the true heat flux is likely to be higher than estimated in the following table: Table 1. FLIR heat flux estimates. λ = 2.7 (constant) bo = 3.2 (constant) W2 = wind speed at 2 m height (m/s) es = vapor pressure of water at Ts (mbar) e2 = vapor pressure of water at 2 m height (mbar) C = 0.61 (constant) Ts = water surface temperature (°C) Ta = air temperature (°C) Tsv = virtual water surface temperature (°C) Tav = virtual air temperature (°C) hs = enthalpy of hot spring water hamb = enthalpy of water at ambient temperature ρw = density of water (kg/m3) 60 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 27 Heat flux estimates are sensitive to wind speeds, as shown in Figure 20. Using a wind speed of 1.5 m/s, the heat flux estimated using FLIR data is 6.96 MW, which corresponds to a flow rate of 0.90 ft3/s, equivalent to 404 gpm. 5.2.4 Discussion Aerial FLIR surveys have been a useful tool in the initial stages of geothermal exploration at PHS. For geothermal exploration using aerial FLIR surveys at systems similar to PHS, 1 to 2 m spatial resolution appears to be sufficient to estimate heat flux using the steps outlined above. A springtime FLIR survey is likely more useful than a fall FLIR survey for identifying blind geothermal resources in high-latitude snow-covered regions where the hot ground contrasts well with cooler snow-covered areas. Combined optical and FLIR airborne surveys offer a relatively inexpensive addition to geothermal resource exploration for targeting further field-based data collection strategies. In logistically challenging areas, such as many areas of Alaska, these surveys may be the most cost-effective method for the first phase of geothermal exploration. While airborne surveys were limited to the use of optical and FLIR cameras, the future use of multispectral or hyperspectral imaging sensors, consisting of several spectral bands in the near- and shortwave-infrared regions, may better characterize the vegetation signatures and alteration minerals associated with the geothermal activity. Heat budget modeling performed in this study estimated that heat flux and flow rates of geothermal waters can be transferred to the characterization of both low-temperature and high- temperature geothermal resources. 6. GEOPHYSICAL SURVEYS In collaboration with the USGS, ACEP conducted geophysical surveys between 2010 and 2013, including a gravity survey in 2010, a high-resolution airborne magnetic and electromagnetic (EM) survey in 2011, and a magnetotellurics (MT) survey in 2012. The goal of these surveys was to provide the regional geophysical framework of the area and help delineate key local and regional structures controlling hydrothermal fluid flow, and characterize the basin geometry and depth to bedrock. Figure 20. The effect of wind speed on heat flux is estimated from fall 2010 and spring 2011 FLIR data for the Pilgrim geothermal system area. 61 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 28 6.1 Gravity Surveys The PGS gravity data were obtained in 1979 and 1980 (Kienle and Lockhart, 1980; Lockhart, 1981) and in 2010 by the USGS. In 1979, 122 stations were occupied along several traverses made on foot and by helicopter, boat, and car. In 1980, one 43 km long north–south regional line was run through PHS (Figure 21). Stations in 1979 were generally along lines, and most stations were 1 to 3 km apart. In the immediate vicinity of the thermal springs, stations were more closely spaced. Station spacing along the 1980 line was anywhere from 1 to 5 km apart. These surveys lacked precise elevation control. Two sets of closely spaced gravity contours trending east–northeast and north–northeast and intersecting a short distance southwest of the thermal area were hypothesized to result from a down dropped basement fault block (Kienle and Lockhart, 1980). The 295 USGS gravity stations in 2010 were located along five north–south lines and one northeast–southeast line, with an additional scattering of more regional points. These data have been merged with the 1979 and 1980 data (Glen et al., 2014). The additional data generally confirmed the earlier contour pattern, with a pronounced gravity low centered about 4 km southwest of the thermal springs being the dominant feature in the valley (Figure 22).The second and more dominant regional gravity feature is along the Kigluaik Mountains range front 2½ miles south of the thermal springs. Figure 21. Gravity stations are labeled on a topographic map of the Pilgrim Hot Springs region. Stations shown by red dots are from the 2010 survey. The earlier stations are shown by gray dots. 62 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 29 6.2 Airborne Magnetic and Electromagnetic Surveys In October 2011, airborne and electromagnetic surveys were flown over the hot springs area. The USGS was primarily responsible for managing this program and interpreting the data. Survey details can be found in Appendix E. About 556 km were flown along north–south lines with east–west tie lines. The mean survey drape of the instrument was 38.2 m. The contractor, Fugro, performed the basic data processing, and the USGS applied additional processing with derivative and filtering methods (Glen et al., 2014). Aeromagnetic data usually provide the most complex and ambiguous geophysical data normally used in geothermal exploration, and the PHS aeromagnetic results live up to this reputation. Glen et al. (2014) note magnetic highs in the vicinity of the PGS and further northwest, and a pronounced magnetic low along the Kigluaik Mountains range front on a reduced-to-pole Figure 22. Isostatic residual gravity map from Glen et al. (2014) used to map the structural basin. Light blue (immediately below Pilgrim Hot Springs) correlates to a basin depth to basement at 320 m (corroborated with drilling contact). Southwest of the hot springs, the deeper basin, indicated by dark blue, is estimated at about 800 m depth. Shallow areas are represented by red. 63 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 30 magnetic field map (Figure 23). A residual reduced-to-pole magnetic map shows a more complicated pattern of shallow-sourced anomalies, with a small magnetic low in the immediate vicinity of the thermal springs (Figure 23). Two narrow northeast–southwest-trending anomalies northwest and north–northeast of the hot springs have magnetic signatures in good alignment with mapped mafic dikes in the Kigluaik Mountains (Glen et al., 2014) and may represent possible dikes that either have not yet been found on the surface or do not quite reach the surface. A magnetic lineation map based on maximum horizontal gradients shows that the PGS is in a somewhat unique position, where two trends terminate as they intersect a third trend. A broad east–west trend is largely terminated by a northeast–southwest trend, and a northwest– southeast trend is terminated by a northeasterly trend (Figure 24). The same generalized trends are present on the isostatic gravity map (Figure 22), giving additional credence to these regional magnetic trends. The depth extent of the electromagnetic survey is in the range of 20 to 125 m (Figure 3 in Glen et al., 2014). The most striking low resistivity in the survey area is centered on the PGS and is approximately co-located with the thawed area at shallow depths near 15 m (Figure 25). A much larger but less intense shallow resistivity anomaly is located north and northeast of the center of the PGS, and overlies the known, but Figure 23. Magnetic field maps from Glen et al. (2014). Magnetic highs appear as reds and pinks, gravity lows as blues and purples, in the reduced-to-pole magnetic anomaly map (left). Magnetic highs appear as reds and pinks, gravity lows as blues and purples, in the differential reduced-to-pole map (right). Figure 24. Magnetic lineations interpreted from maximum horizontal gradients of pseudogravity. Colored by trend (EW, red; NW, blue; NE, green). From Glen et al. (2014). 64 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 31 largely unexplored, northeastern thermal anomaly. At a slightly greater depth of 35 m, the northeastern thermal anomaly area is the dominant low-resistivity area, and the core of the known PGS is no longer particularly low in resistivity (Figure 25). The presumed upflow area is in this region and will be discussed later in this report. Two areas west–southwest and southwest of the PGS have interesting low-resistivity values at a depth of 15 m, and it is speculated that they result from graphitic metamorphic rocks (Glen et al., 2014). The higher-resistivity rocks reflect metamorphic bedrock and coarser-grained glacial outwash sediments. Unfortunately, the electromagnetic survey was not capable of penetrating to depths near the PGS, which would have helped locate the upwelling zone, but the survey does offer the possibility that other and possibly even larger thermal areas are in the Pilgrim Valley. 6.3 Magnetotellurics Survey In August 2012, Fugro obtained 59 magnetotellurics (MT) soundings at the PGS. Spacing between sites varied from about 300 feet to about 1800 feet, with less-dense coverage away from the center of the known thermal anomaly (Figure 26). The outer ring of MT sites was specifically chosen to extend beyond the known limits of the shallow thermal anomaly in all directions except toward the northeast. No sites were occupied north of the Pilgrim River, as the intent was to locate the upwelling of the PGS, not to study the more inaccessible northeastern thermal anomaly. At a depth of 25 m, approximately the known depth of the shallow thermal aquifer, the MT data show a nearly circular low-resistivity area about 900 m in diameter, with resistivities as low as 2 Ohm-m. The MT is probably responding to the high salinity of the PHS thermal fluid. The area with resistivity less than about 5 Ohm-m is a good approximation of the 49°C (120°F) temperature contour defining the shallow thermal aquifer. Above about 5 Ohm-m, the resistivity contours are tight, rapidly climbing to values above 100 Ohm-m. The high resistivity values probably reflect the low-salinity permafrost surrounding the thermal anomaly. Figure 25. Airborne EM resistivity slices shown at 15 m (left) and at 35 m (right). From Glen et al. (2014). 65 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 32 The sharp margins of the MT anomaly are also clearly defined in cross-sectional view (Figure 27). The temperature contours at shallow depths generally behave in a sharply bounded fashion, near the edge of the shallow thermal aquifer (Figure 28). The obvious exception to the sharp boundaries for both datasets is toward the northeast, where the shallow thermal aquifer has its greatest known length. The MT anomaly also extends that direction. The MT data show a short “nose” extending southwest of the PS-5 hole that was not picked up by the temperature data. At a depth of 50 m, a short distance below the shallow thermal aquifer the resistivity increased slightly, but the circular anomaly core is still present (Figure 28). Figure 26. Magnetotellurics site locations. Figure 27. Resistivity at Profile D from a 1D MT inversion. 66 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 33 The thermal minimum in the deeper PGS wells occurs near a depth of 100 m. At this depth, the MT shows the smallest areal extent of less than 3 Ohm-m resistivity of any of the depths (Figure 29). The lowest resistivity values are now centered about 85 m southeast of well PS-12-2. At depths between 100 and 300 m, which are within the Quaternary alluvium, the area of less than 5 Ohm-m gradually expands and moves toward the northwest (Figure 29). Between depths of 300 and 500 m, in the metamorphic bedrock, the lowest resistivity values shift noticeably about 0.5 km to the southwest; by 500 m, they are centered beneath well PS-5 (Figure 29 and Figure 30). By a depth of 1000 m, the lowest resistivity values have radically shifted east and northwest of the shallow thermal anomaly (Figure 30). Since the MT survey was run to help locate the upwelling zone, the question of whether this survey was successful must be addressed. While MT clearly succeeded in locating and outlining the shallow thermal anomaly, there is no clear evidence that MT located the thermal upwelling. The small volume of low-resistivity values near PS-12-2 at depths of 100 to 200 m within the alluvium cannot directly represent thermal upwelling, given the temperature profile of PS-12-2. The large horizontal shifts of low-resistivity areas below 300 m in metamorphic basement rocks may represent some larger-volume conductor(s) other than hot water. The full Fugro (2012) report, which discusses this topic in detail, is included as Appendix L. Figure 28. Resistivity maps at 25 m and 50 m from the blind 3D MT inversion. The red line represents the 49°C (120°F) temperature contour in the shallow thermal aquifer. The straight black lines are the MT transects. 67 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 34 Figure 29. Resistivity maps at 100 m, 150 m, 200 m, and 300 m depths from the blind 3D MT inversion. The red line represents the 49°C (120°F) temperature contour in the shallow thermal aquifer. 68 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 35 7. DRILLING ACTIVITIES The bulk of the money and effort invested in the geothermal exploration project at PHS was used for drilling activities, with the aim of obtaining accurate subsurface temperature data and identifying the main upwelling zone. Drilling ranged from very shallow activities carried out with a simple gasoline-powered backpack drill, to large-diameter drilling that required a large rotary drill rig and mud circulation systems capable of drilling a 14-inch-diameter well to bedrock. Additionally, the valves on the wellheads of the wells drilled in 1979 and 1982 were replaced to cease uncontrolled artesian flows and allow the wells to be logged in a static state. The deep holes and wells that have been drilled at the site since 1979 are shown on Figure 4. Figure 30. Resistivity maps at 400 m, 500 m, 750 m, and 1000 m depths from the blind 3D MT inversion. The red line represents the 49°C (120°F) temperature contour in the shallow thermal aquifer. 69 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 36 7.1 Permitting Before any of the drilling activities described in this report could occur, a variety of permits and land usage agreements had to be secured. Permits were obtained in several phases as drilling plans were refined and new data were input into geothermal models of the area. Land use permits were also obtained so that UAF and its contractors could legally perform activities associated with geothermal exploration on the landholdings of various entities. The land use agreements, permits, and waivers that were obtained for this project are summarized in Table 2. Table 2. Permits and approvals 7.2 Legacy Wellhead Repairs During 1979 and 1982, six wells penetrating the shallow thermal aquifer were drilled to depths of 1000 feet. These wells were never plugged and were abandoned. Due to a lack of maintenance, the wellheads were in extremely poor condition when examined by ACEP in 2009. The wellheads had to be repaired to control artesian flows and permit new static temperature logs and water samples. During an initial site visit in July of 2010, an assessment of each well was made and work plans were developed. Wellhead repairs occurred September 13–18, 2010. The team completing the repairs was able to replace the master gate valves on wells PS-1, PS-3, PS-4, and MI-1. Wells PS-2 and PS-5 were not found to be leaking, and the team was not able to replace the gate valves because of swampy conditions around the wells, which restricted heavy equipment access. At each of the four repaired wellheads, the team removed the existing gate valves while pumping down the water and installed new stainless steel valves. Detailed repair descriptions for each well are given in Appendix D. Images from the repair of well PS-4 are shown in Figure 31 and Figure 32. Entity Permit or Approval Alaska Oil and Gas Conservation Commission Permits to drill Alaska Department of Environmental Conservation Storage/discharge of drilling waste solids, Waste water discharge approval/ waiver National Environmental Policy Act Project review Department of Natural Resources Temporary water use permit for drilling makeup water and flow testing Alaska Department of Fish and Game Project approval, Waiver of fish habitat permit for flow test U.S. Bureau of Land Management Permit for road or trail improvements, Gravel pit use U.S .Army Corp of Engineers Verification that project is in Nationwide Permit 6 accordance Mary’s Igloo Native Corporation Land use permit Unaatuq, LLC Exploration license Bering Straits Native Corporation Exploration license State Historic Preservation Office Finding of no historic properties effected 70 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 37 Multiple static and flowing temperature and pressure logs were obtained for all wells except for PS-2, where the wellhead has sunk into the soft ground and the master valve is inoperable. Water samples were collected from these wells and from the natural hot springs for chemical analysis. 7.3 Shallow Temperature Survey At shallow depths, the PGS is dominated by a strong lateral flow of geothermal water, identified three decades ago when the first six wells were drilled into the system. The maximum temperature of this shallow aquifer is slightly below boiling, and the depth to the most hydraulically conductive part of the aquifer is less than 100 feet. This combination of factors produces very high shallow- temperature gradients above the thermal aquifer and sharp temperature declines below the aquifer. The smooth nature of the six early shallow temperature profiles strongly suggests that the aquifer began transmitting hot water in the relatively recent past and that the lower temperatures beneath the aquifer are a result of downward conduction of heat from the aquifer—not a flow of cold water beneath the thermal aquifer. If there were a counterflow of cold water, complexity such as isothermal segments in the temperature profile separated by short intervals of extremely high temperature gradients would be expected. This combination of characteristics at PHS allows the possibility of defining the shallow thermal aquifer with abnormally shallow holes compared with most other geothermal systems. Characterizing the shallow thermal aquifer allows definition of the directions of thermal fluid flow within the aquifer and recognition of the hottest part, which most likely would overlie thermal upwelling beneath the aquifer. The absence of bedrock and coarse conglomerate in the vicinity of PHS is also an important factor that allowed consideration of low-cost and unconventional drilling techniques. The flat, Figure 31. Areas of leaking, scale, and corrosion are shown on PS-4. Figure 32. The PS-4 completed replacement valve installation. 71 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 38 swampy topography at PHS is an advantage to the extent that it minimizes topographic effects at shallow depths; however, it also inhibits access to much of the area with machinery. The first shallow temperature holes at PHS were installed in 1979, when about 70 “pipes” were hand driven to a maximum depth of 5 to 9.5 m (Harrison and Hawkins, 1979; Osterkamp et al., 1980). An isothermal map at a depth of 4.5 m was prepared, outlining the central part of the shallow thermal anomaly with temperatures between 30°C and 80°C (86°F–176°F). Effort was focused on the heart of the shallow thermal anomaly, and none of the holes was deep enough to penetrate into or beneath the shallow thermal aquifer. 7.3.1 Backpack Drilling Program Two additional shallow temperature surveys were attempted in April 2011 using a small backpack-mounted drill. Thirty-one holes were drilled to depths of 3 m while the area was still snow-covered and could be accessed by snowmobile. However, a number of challenges arose including holes collapsing before tubing could be installed and snow depths of 2 m. These challenges limited the ability to install as many holes as desired to a uniform depth, which presents difficulty with interpretation. The backpack-drilling effort did not produce results much better than the effort made in the 1970s. 7.3.2 Geoprobe Drilling Program Discussion with USGS project partners revealed that they had a self-contained track-mounted direct-drive Geoprobe unit, touted as capable of driving pipes to depths of less than 30 m. The unit is highly mobile, and at a weight of 5000 pounds, was light enough to travel on trails in the PHS area with minimal impact. The Geoprobe unit drives small-diameter sealed pipes into the ground without the need for circulatory fluids (Figure 33), eliminating the mud system that traditional drill rigs require. During the summer of 2011, sixteen Geoprobe holes with an outer pipe diameter of 2.25 inches (5.7 cm) and a hole diameter of 1.5 inches (3.81 cm) were installed to a maximum depth of 109 Figure 33. Installing Geoprobe holes at Pilgrim Hot Springs. 72 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 39 feet. Locations and temperatures are shown in Figure 34. Prior to the end of the 2011 season, all holes were decommissioned by pulling the pipes and sealing the holes with grout as the pipes were removed. Figure 34. Location of Geoprobe holes and their temperatures in Fahrenheit at 60 feet. 73 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 40 Nearly all of the Geoprobe holes from 2011 failed to reach beyond 80 feet deep; none penetrated into or beneath the shallow thermal aquifer. In 2012, smaller pipe (1.25 in. outer diameter; 0.5 in. inner diameter) was used in 54 holes, enabling deeper penetration (Figure 34). The deepest hole reached 154 feet. The majority of these holes still have positive temperature gradients, but some have encountered isothermal conditions indicative of having penetrated the shallow thermal aquifer, documenting its maximum temperature (Figure 35). Phase 2 drilling produced known depths of the aquifer, so it was possible to extrapolate some of the Geoprobe hole temperature profiles and better define the flow pattern within the shallow thermal aquifer. All Geoprobe locations and depths are shown in Appendix C. 7.4 Deep Drilling Deep drilling at PHS occurred over three different field seasons: 2011, 2012, and 2013. Holes and wells drilled more than 500 feet in total depth required permits from the Alaska Oil and Gas Conservation Commission, while those shallower than 500 feet did not. During 2012, a blowout preventer (BOP) was required when drilling below 1000 feet. In 2013, a waiver was obtained, and a BOP was not required. Drilling in 2011 and 2012 was accomplished with USGS Alaska Rural Energy Project equipment and personnel. This drilling was done with an Atlas Copco CS- Figure 35. The temperature logs from all Geoprobe holes show a wide variety of temperatures and profile shapes. A shallow thermal aquifer of varying depths and characteristics is clearly visible. 0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 20 40 60 80 100 120 140 160 180 200 Depth (feet) Temperature (F) Pilgrim Hot Springs Geoprobe Temperature Logs 74 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 41 1000-P6L drill rig, while in 2013, drilling was done by MW Drilling using their Schramm Model T555 Rotadrill rig based out of Anchorage, Alaska. Well schematics and detailed descriptions for each well are found in Appendix A. In June 2011, prior to drilling activities, researchers from ACEP and the USGS conducted an aerial inspection of PHS via helicopter to locate suitable temperature gradient (TG) hole drilling locations. Nine possible drill sites were identified. The initial drilling targets were northeast of the historic hot springs, located on land owned by the Mary’s Igloo Native Corporation. This decision was based on data from the 1982 drilling effort, especially the cool temperatures and low bottom-hole gradient measured in well PS-5 (Figure 5). Drilled in 1982, PS-5 was the deepest well that had been drilled during that effort; it also recorded the coolest temperatures. This information clearly showed that the upwelling zone could not be located south of the existing well field. The upwelling zone appeared to be located northeast of the existing wells, with evidence for this supported by the appearance of a thawed zone in the northeast area. In 2011, drilling was sited as far north as logistically possible, where the USGS drill rig could access the area using existing primitive roads and trails. Drilling took place just southwest of the Pilgrim River and resulted in TG holes S-1 and S-9 (Figure 4). The temperatures measured in these holes were significantly cooler than the temperatures measured in the existing wells, suggesting that these two holes were too far to the northeast and that the upwelling zone must be closer to the historic hot springs. In 2012, three TG slim holes were drilled on the PHS property owned by Unaatuq, LLC. Drilling logs for these wells are shown in Appendix J. The 2012 drilling activities moved farther to the south, with the first hole (PS-12-1) located slightly north of the historic orphanage buildings and the second and third holes (PS-12-2 and PS-12-3) drilled near the existing well field. The equilibrated temperature profiles (shown in Appendix B) collected from these three wells show temperature reversals beneath the shallow thermal aquifer, indicating that they are not directly over the upwelling zone. The holes drilled in 2011 and 2012 used sealed casing cemented in place. The holes were only permitted as TG holes and were not intended to have the ability to access fluids in the geothermal aquifer. Once drilling was completed, the casing was filled with water so that a temperature probe could be lowered into the hole to record accurate static aquifer temperature profiles. Wells drilled in 2012 were at first assigned names by the Alaska Oil and Gas Conservation Commission: TG-1, PS-12-3, and PS-12-9. In order to be consistent with the existing nomenclature, the names of these wells were changed so that TG-1 became PS-12-1, PS- 12-3 became PS-12-2, and PS-12-9 became PS-12-3. In 2013, an attempt was made to drill a large-diameter well into the predicted upwelling zone and test fluid production from the aquifer above the bedrock. The drilling methods used were similar to those used in 2011 and 2012, but makeup water was pumped from the slough on the property in accordance with state and federal regulations. The first drill site was chosen based on data from the wells drilled in 2012, which suggested that those wells surrounded the upwelling. When the first well drilled in 2013—PS-13-1—encountered the usual large temperature reversal and showed lower temperatures than hoped, it was completed in the shallow aquifer. Two more 75 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 42 small-diameter wells were drilled in 2013 to 400 feet. All wells drilled in 2013 showed temperature reversals, indicating that they were not directly over the upwelling. All three holes drilled in 2013 used perforated casing or well screen and have artesian flows. Details about each deep TG hole and well are described in Appendix A. 8. WATER CHEMISTRY The geochemistry of the thermal fluid at PHS is one of the primary reasons why so much effort over so many years has been put into exploring this geothermal system. In addition to its relatively hot surface temperature, thermal fluid at PHS has the highest predicted quartz geothermometer temperature (137°C) and one of the highest Na-K-Ca geothermometer temperatures (146°C) of the thermal springs on the mainland of Alaska (Miller et al., 1975). Chemical analyses of PHS thermal waters now cover a century (Waring, 1917; Miller et al., 1975; Liss and Motyka, 1994; Benoit et al., 2014b), and an extensive water chemistry database has been assembled (Table 3). The thermal water at PHS is relatively high in sodium and chloride, but stable isotope analyses of thermal and cold waters at this location show that the thermal water is derived from local meteoric runoff and not from seawater (Miller et al., 1975; Liss and Motyka, 1994). Only three noncondensible gas samples from PHS thermal waters have been analyzed, and these contain mostly methane and nitrogen and are relatively high in hydrogen (Liss and Motyka, 1994). Gas geothermometry results indicate subsurface temperatures from 113°C to 230°C (235°F–446°F). The surface flowing temperatures and brine chemistry of some of the PHS wells have changed with time (Liss and Motyka, 1994), but these changes have since been shown to result from changes in fluid entry points in those wells (Benoit et al., 2014b). Table 3. Pilgrim Hot Springs well chemistry in PPM Sample Date T°C pH Na K Ca Mg Li B SiO2 HCO3 CO3 SO4 CL F Spring 1915 70 1590 61 545 7.4 87 21 25 3450 Spring 1972 82 6.75 1450 61 530 1.4 4.0 2.4 100 30.1 24 3346 4.7 Spring 1982 55 6.8 1660 59 542 1.0 4.5 2.2 91 36 15 3360 4.3 Spring 1993 42 6.5 1580 65 569 1.5 4.0 2.7 86 19 18 3530 4.7 Spring 1993 55 6.8 1660 59 542 1.0 4.5 2.2 91 36 15 3360 4.3 Spring 2012 6.65 1480 62.8 508 0.38 3.6 2.0 86 14 22 3350 4.6 Spring 2014 73 6.63 1400 58 460 1.00 3.4 2.1 80 15 3500 PS-1 1979 90.5 6.4 1828 75 518 0.9 3.9 2.5 95 16 16 3590 4.8 PS-1 1982 92 7.5 1720 60 511 0.9 4.7 2.3 94 30 19 3420 4.4 PS-1 1993 82 7.1 1560 65 545 0.6 4.2 2.4 90 20 7 3460 5.3 PS-1 2010 79 7.1 1530 61.6 519 1.21 3.5 2.2 83 27.8 14.3 3460 4.5 PS-2 1979 90 6.4 1820 75 516 0.9 3.9 2.3 101 19 15 3540 4.8 PS-2 1982 96 7.3 1510 57 516 0.9 4.7 2.3 92 26 19 3420 4.5 PS-3 1982 75 8.0 592 25 260 0.4 2.0 1.0 60 36 15 1430 1.3 PS-3 1993 65 6.8 1100 43 441 0.6 3.2 1.5 67 27 6 2450 2.9 PS-3 2010 67 7.0 1140 40.9 412 0.85 2.8 1.7 71 23.7 10.8 2650 3.0 PS-4 1982 48 8.6 115 4.8 23 0 0.3 0.5 35 80 11 284 0.5 PS-4 1993 45 8.6 146 7.8 98 0.2 0.2 0.2 27 48 1 386 0.3 PS-4 2010 44 8.52 152 5.9 73 0.14 0.5 28 34 9.4 353 0.4 PS-4 2013 44.6 8.47 128 5.5 45 0 0.4 0.2 29 39.9 1.5 9.3 260 0.6 PS-5 1993 32 9.6 36 1.1 2 0.2 0.1 0.6 21 81 5 6 0.5 PS-5 2010 30 9.6 36 1.09 1 0 0.1 20 49.6 5.4 2 0.5 MI-1 1982 22 9.7 16 0.5 5 0 0.1 21 37 9 5 0.2 MI-1 1993 31 8.3 29 1.5 23 0.6 0.2 0.1 20 32 10 66 0.2 MI-1 2010 29 7.8 130 4.4 93 0 0.5 21 25.8 9.5 337 0.2 PS 12-3 2012 65.5 7.52 731 29.9 281 0.78 1.8 1.0 51 30.6 8.2 1640 1.9 PS 13-1 (open to 1036 ft) 2013 70.5 7.51 537 26.1 236 0.4 1.4 0.8 54 25.1 9.3 1300 1.4 PS13-1 2013 77 7.27 1090 50.9 370 0.7 2.6 1.5 79 22.8 12.4 2500 3.3 76 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 43 Sample Date T°C pH Na K Ca Mg Li B SiO2 HCO3 CO3 SO4 CL F (shallow Completion) PS 13-1 300 gpm 2014 79 7.26 1000 35.0 250 2.0 1.5 59 18 2500 PS 13-1 60 gpm 2014 77 7.05 950 37.0 250 2.1 1.4 67 15 2300 PS 13-2 2013 71 8.95 124 25 49 0 0.3 0.2 62 39.4 11.1 5.8 265 0.5 PS 13-2 55 gpm 2014 69 7.52 53 3.1 9 0.2 0.1 54 62 5.5 65 PS 13-3 2013 79 7.27 1070 46.3 373 0.7 2.5 1.4 74 22.1 12.3 2424 3.0 PS 13-3 60 gpm 2014 78 6.97 920 37.0 280 2.2 1.3 66 16 2200 There is an obvious mixing trend between dilute cold groundwater, and primitive geothermal fluid, exemplified by sodium and chloride contents (Figure 36) and temperature (Figure 37) (Liss and Motyka, 1994). The same relationship is shown in cross plots involving all other chemical species except sulfate, which has more scatter. Flowing temperature profiles show little or no mixing of different fluids within the wellbores (Benoit et al., 2014b) and thin discrete aquifers with discrete chemistries that are supplying the PHS wells. The mixing trend shown in Figure 36 does not fall along the line of charge balance where sodium ions equal the chloride ions. The thermal fluid at PHS is deficient in sodium, and this deficiency is balanced by an abundance of calcium, causing the apparent mixing to diverge from the Na-Cl line. The calcium content of the primitive geothermal fluid is greater than 525 ppm, and exceptionally high when compared with most other low-salinity geothermal waters throughout the world. A major disappointment of recent exploration at PHS has been the inability to find the more optimistic temperatures predicted by geothermometry. The exceptionally high gas geothermometry values have always been viewed as questionable (Liss and Motyka, 1994), but the low magnesium content and the neutral chloride nature of the thermal fluid along with the Figure 36. The mixing trend between sodium and chloride is shown for all samples collected from the Pilgrim Hot Springs site. 0 500 1000 1500 2000 2500 3000 3500 4000 0 500 1000 1500 2000Chloride ppm Sodium ppm Pilgrim Springs Chemistry Hot Springs PS-1 PS-2 PS-3 PS-4 PS-5 MI-1 Lake NaCl equivalent line 77 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 44 quartz and the Na-K-1/3Ca geothermometry appeared to credibly predict temperatures of 130 to 145°C (266°F–293°F). To date, all drilling to depths as great as 1294 feet has resulted in a measured maximum temperature of only 91°C (196°F). To compound the frustration of this finding, temperature profiles in the deeper wells cannot be extrapolated to significantly greater depths to predict reliably where these higher temperatures may be present. Of course, it is always possible that higher temperatures are located at a much greater depth or lateral separation. However, other geothermometers might be more appropriate for PHS. The chalcedony geothermometer predicts subsurface temperatures of 99°C to 111°C (210°F–232°F), and the Na- K-4/3Ca geothermometer gives values near 120°C (248°F). The 120°C value is still 29°C above the maximum measured temperature and begs the question of whether the geothermometer is valid for the PGS brine chemistry. The 525 ppm of calcium in the PGS water is an obvious suspect in raising the question of whether thermal waters with exceptionally high calcium content provide accurate geothermometry calculations. 9. FLOW AND INTERFERENCE TESTING The first flow testing and interference testing of the PGS were performed in 1982, when well PS- 1 was flowed at 30 to 35 gpm and pressures were recorded in well PS-2. Type-curve matching of the drawdown gave an estimated permeability of 4.5 darcys (Economides et al., 1982). In 1982, the productivity of the wells ranged from 2.5 to 19 gpm/ft, and the transmissivity of the wells ranged from 300 to 40,000 gpd/ft (Kunze and Lofgren, 1982). Figure 37. Chloride content is shown along with well temperature. The PS-13-2 chloride content appears to be low, given its temperature. 0 10 20 30 40 50 60 70 80 90 100 0 1000 2000 3000 4000Sampling Temperature C Chloride ppm Pilgrim Springs Sampling Temperature versus Chloride Content Hot Springs PS-1 PS-2 PS-3 PS-4 PS-5 MI-1 13-1 13-2 ? PS 13-3 PS 12-3 78 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 45 9.1 Interference Testing of Wells PS-3, PS-4, and MI-1 Three interference tests using downhole temperature and pressure monitoring were performed in September 2013 (Benoit, 2013). The first test involved “static” pressure and temperature monitoring of well PS-3, with wells MI-1 and PS-4 being flowed with different start and stop times over a period of two and a quarter days. The second test involved flowing well PS-4 for 3 hours and monitoring wells PS-1 and PS-5. The third test was a mirror image of the first test, with well PS-3 being flowed for 3.5 hours and downhole pressure and temperature monitoring in wells PS-4 and MI-1. The interference tests conducted on September 7, 8, 9, 11, and 22 generally confirmed the observations made by Woodward-Clyde during their flow tests in 1982. Examples of the temperature and pressure responses during these tests are shown as Figure 38 and Figure 39. Wells PS-1 and PS-2, completed in the shallow thermal aquifer, do not quickly or obviously communicate with the deeper and cooler aquifers exposed in wells PS-3, 4, 5 and MI-1. More precise tools available in 2013 have shown that wells MI-1, PS-3, and PS-4 have a rapid but barely detectable pressure communication of 0.1 to 0.25 psi. This communication occurs at flow rates of 50 to 100 gpm from individual wells. This small pressure communication creates a much stronger and surprising temperature change in the static well PS-3 when wells MI-1 and PS-4 are flowed. The speed with which this temperature communication occurs indicates that the small changes in pressure create flow rate changes that quickly change the water flow past the tool in the “static” PS-3 well. More likely, the temperature changes are related to flow rates and mixing than to a single fluid entry changing its temperature. No obvious temperature changes appear in well PS-4 when PS-3 is flowed. Well MI-1 showed some small temperature changes when PS-3 was flowed, but these changes are not Figure 38. PS-3 downhole pressure during interference testing. 84 84.5 85 85.5 86 86.5 0 5 10 15 20 25 30 35 40 45 50 55Pressure at 59.35 Meters Below Flange (psig) Time (hours) PS-3 Downhole Pressure Record Sept. 7-9, 2013 Downhole Pressure Open MI-1 Reinstall pressure tool Shutin MI-1 Open PS-4 Shutin PS-4 79 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 46 as clearly related to starting and stopping PS-3 flow. Wells PS-3, PS-4, and MI-1 have relatively similar depth permeable intervals, which is a first-cut explanation for their measurable short-term communication. Some background temperature and pressure trends in the PHS wells are not understood and will require additional and/or longer-term monitoring to understand. The full flow-testing report for September 2013 is shown in Appendix H. 9.2 Interference Testing of PS-3, PS-13-1, and PS-13-3 Interference testing and flow testing of the 2013 wells were conducted twice about 6 months apart. The first testing occurred in February 2014 during a winter trip to the hot springs. The purpose of this trip was primarily for collecting equilibrated temperature logs of the 2013 wells, and time available for flow testing was limited. The temperature and pressure were monitored in wells PS-3, PS-13-1, and PS-13-3, while well PS-13-1 and later well PS-13-3 were allowed to flow at natural artesian rates of 50 to 70 gpm (Appendix I). Flows were visually estimated, as no flow metering was available. Well PS-13-3 was allowed to flow for just under 5 hours, and immediately after the flow was cut off, well PS-13-1 was opened and allowed to flow overnight for 12 hours. At artesian flow rates, the recorded pressure and temperature effects between the wells were extremely minimal, on the order of 0.2 psi and 0.02°C. In each well, productivity was approximately 20 gal/psi. Productivity will be discussed in further detail in the next section. 9.3 Flow Testing of PS-13-1 To date, the most significant flow test at the PGS was conducted between September 15 and 17, 2014, by airlifting well PS-13-1. The airlift was accomplished using thin-wall 1-inch-diameter aluminum tubing with a dispersion head on the bottom and an Atlas Copco trailer-mounted air compressor rated at 100 psi and 185 cfm. This hardware was supplied by Howard Trott of Potelco and rented locally in Nome. A 6-inch Krohne magnetic flow meter (magmeter), supplied by ACEP, was used to measure the flow rates. The surface equipment is shown in Figure 40. The Figure 39. PS-3 temperature response during 2013 interference testing. 75 75.5 76 76.5 77 77.5 78 78.5 79 0 5 10 15 20 25 30 35 40 45 50 55Temperature C Time (hours) PS-3 Downhole Temperature Record Sept. 7-9, 2013 Downhole Temperature Open MI-1 Reinstall pressure tool Shutin MI-1 Open PS-4 Shutin PS-4 80 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 47 5-inch-diameter dispersion head was backed up with 1/8 inch aircraft cable to prevent accidental loss of the downhole equipment in the well. The air-water mixture flowed with considerable turbulence into the first tank shown in Figure 40. As the water flowed into the second tank, no turbulence occurred, and the tank provided adequate head to push the water through the magmeter and out a 240-foot-long 6-inch PVC pipeline to flow through a hot springs pond, where the water cooled. The first airlift only lasted about an hour, as the flow was limited by a constriction in the flow line downstream of the second tank. Expansion of the flow line caused a short flexible hose to partially collapse, reducing the flow rate out of the second tank. This first test was more a test of the equipment than a test of the resource. The aluminum tubing was run to a depth of 12.2 m below the top of the standpipe (8.8 m below ground level). The average airlift flow rate during the first test was 172 gpm, and the magmeter readings were confirmed by measuring that it took 18 seconds to fill a 55-gallon drum from the discharge of the pipeline into the flow through a hot springs pond. Pumping at a higher air rate increased the flow rate to 177 gpm, but resulted in the water overflowing the top of the wellhead standpipe. While the well was being airlifted and the wellhead appeared to be stable, a Kuster tool was run to a depth of 30 m below the top of the master gate, with the heavier aircraft cable used as a backup in case the small 1/16-inch-diameter cable normally used on the reel was cut. The Kuster tool hung in the well just over a half hour before the air was cut off. Once the air was cut off, the well resumed its natural artesian flow of 55 gpm, and the Kuster tool remained hanging in the well overnight to record pressure buildup. There was no wing valve on the flow line to allow the well to be shut in Figure 40. Surface equipment used for the airlift of PS-13-1. The magmeter is in the silver spool between the black and white parts of the flow line. The black large-diameter hose serving as a standpipe on top of the wellhead was needed to prevent water from overflowing the top of the wellhead, which could not be sealed. The clamp holding the aluminum tubing is visible on top of the standpipe. The blue hoses are the air lines coming from the air compressor. 81 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 48 with the aluminum tubing and Kuster tool hanging inside it. On the morning of September 16, additional flow line parts were obtained in Nome, and the line was modified to remove the constriction. The aluminum tubing in the well was deepened from 12.2 m to 21.9 m below the top of the standpipe on the wellhead (18.6 m below ground level). This depth was about the maximum practical for one person on a large A-frame ladder to raise and lower the downhole equipment; it was also the maximum depth at which the on-hand larger- diameter aircraft cable could be used to hold and protect the Kuster tool. In the second test, there was adequate confidence to run the Kuster tool into the well under artesian flow conditions before starting the airlift. The air volume was quickly increased in three steps to find the maximum airlift rate that would not flow water out the top of the wellhead. This flow rate was about 300 gpm. The highest flow rate reported briefly by the magmeter was about 350 gpm. The 300 gpm airlifted flow rate was held for about 7.5 hours, until the air compressor was almost out of diesel fuel at 01:00 hours on September 17. After the compressor was shut off, the well continued artesian flow until after the Kuster tool was retrieved late in the morning on September 17. Also on that morning, the downhole hardware was pulled out of the well. Airlifting increased the scatter in the pressure and temperature data as compared with the unassisted artesian flow (Figure 41 and Figure 42). During the first airlift, it is unclear if there was any decline trend in the downhole pressure. The first 15 minutes of downhole data indicate a decline, but perhaps this was simply the tool equilibrating to the downhole conditions (Figure 42). During the second 15 minutes, no decline is evident. At 19:00, the amount of air being pumped was increased for 2 minutes to assess the plumbing system at higher flow rates and was then shut off (Figure 42). The amount flowing through the meter increased by only about 5 gpm to 177 gpm, but water occasionally geysered at the top of the wellhead. A constriction in the soft 6-inch hose between the two tanks limited the flow through the meter. The downhole flowing temperatures were measured below the air injection depth and, therefore, were not cooled by the air injection, as the surface-measured temperatures were. The maximum downhole temperature measured during the first airlift was 78.28°C (Figure 42). Immediately upon shutting off the air, the temperature took a 0.2°C decline and then quickly climbed for the next 13 minutes to its maximum value of 78.8°C, then quickly cooled. The temperature was down to 77°C when the tool was removed the following morning and showed a range of 1.7°C during this logging. During the airlift, the temperature increased slightly. After airlifting ceased, the bulk of the temperature change occurred, first with a short 0.2°C decrease, probably related to the short increased volume airlift, and then with a 0.8°C increase followed by a long decline until the temperature was about 1°C lower than during the airlifting. During this decline, the well was flowing under natural artesian conditions. This variation of temperatures with flow rates demonstrates that there is more than one feed zone for this well, with differing temperatures. Higher temperatures coincide with higher flow rates. A similar response was seen upon stopping the second airlift (Figure 43); however, this response lacked the sharp initial drop in temperature as seen at the end of the first airlift (Figure 42). The maximum temperature recorded after stopping the second airlift was 79.8°C, or 1.0°C hotter than seen after the first airlift stopped (Figure 41). After the second airlift was finished, the 82 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 49 artesian flow temperature declined to 76.9°C, about 0.25°C cooler than that seen after the end of the first airlift. Figure 42. PS-13-1 downhole pressure and temperature record just before and after stopping the first airlift at 19:02 hours on September 15, 2014. 77 77.2 77.4 77.6 77.8 78 78.2 78.4 78.6 78.8 79 37 38 39 40 41 42 43 44 9/15/2014 18:009/15/2014 18:159/15/2014 18:309/15/2014 18:459/15/2014 19:009/15/2014 19:159/15/2014 19:309/15/2014 19:459/15/2014 20:00Temperature C Pressure (psi) Date and Time PS-13-1 Monitoring at 30 m During and After First Airlift Pressure Temperature Increase air injection rate Stop First Airlift Figure 41. Downhole pressure and temperature record of PS-13-1 during the two airlifts. 76 76.5 77 77.5 78 78.5 79 79.5 80 36 38 40 42 44 46 48 50 9/15/2014 18:009/16/2014 0:009/16/2014 6:009/16/2014 12:009/16/2014 18:009/17/2014 0:009/17/2014 6:009/17/2014 12:00Temperature C Pressure (psi) Date and Time PS 13-1 Monitoring at +30 m During Airlifts September 15-17, 2014 Pressure Temperature stop airlift and start 300 gpm airlift stop 172 gpm airlift 83 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 50 This temperature variability points to a fairly complex interplay between two or more feed zones with differing temperatures. This idea led to a detailed flowing log run on the morning of September 18, before the well was shut in and the artesian flow was stopped. This artesian flowing log and a static log run on September 7, 2014, show some of the details of the fluid entry points (Figure 44). The flowing SMU log shows multiple sharp reversals in temperature gradient between depths of 56 and 67 m, which define all the possible fluid entry points. The top of the screen in the wells is at 57.3 m, which is in good agreement with the flowing temperature log. Due to minimal divergence between the flowing and static logs between depths of 65 and 67 m, any fluid entry point in that interval is suspect, as the temperature readings were not stable in that and shallower intervals. The deepest significant fluid entry is at a depth near 65 m, and the shallowest major entry as defined by temperature is near 60 m. All of the defined fluid-entry temperatures are between 77.02°C and 77.18°C on the flowing log. However, the static temperatures in this interval range from 77.4°C to 77.7°C. During airlifting, temperatures as high as 78.25°C to 79.3°C were measured, which had to have come from shallower depths in the well, perhaps as shallow as 35 or 40 m. This fluid would then have had to flow down the outside of the uncemented 14-inch casing and enter the screened interval between 57.3 and 72.5 m (see Appendix A for well schematic). The maximum temperature measured during the airlifting operations was the 79.8°C spike shortly after ceasing the airlift. This temperature is only 0.18°C hotter than the maximum measured tempeature of 79.62°C during the static log prior to flowing the well. Thus, we now have a good idea as to the origin of the fluid producing the temperature spike. Figure 43. Downhole pressure and temperature at the end of the second airlift at 01:00 hours on September 17, 2014. 76.5 77 77.5 78 78.5 79 79.5 80 36 38 40 42 44 46 48 50 9/17/2014 0:009/17/2014 0:309/17/2014 1:009/17/2014 1:309/17/2014 2:00Temperature C Pressure (psi) Date and Time PS-13-1 Monitoring at +31 m Depth at End of Second Airlift Pressure Temperature stop air lift and resumeartesian flow 84 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 51 It was decided that during the airlift the internal wellbore conditions would probably be too severe for the small SMU tool and its delicate electrical cable. The primary use of the Kuster tool was for pressure monitoring, so it was not moved during the airlifting. However, during any future airlift, a traversing Kuster survey should be run. Four major flow rate changes were monitored with downhole pressure changes in PS-13-1 during September 2014. The first was done on September 15, prior to the airlifting, and involved opening up the well so that it could artesian flow. During this flow, the rate was somewhere between 60 and 75 gpm, as measured with a 5-gallon bucket. Three major flow rate changes were then monitored during airlifting while the Kuster tool was downhole (Figure 41, Table 4) and the magmeter was providing the flow rate data. The first flow rate change was the cessation of the first airlift, the second was the start of the second airlift, and the third was the end of the second airlift. All of these changes had natural artesian flow either before or after. None of the changes involved the larger change of going from a static condition to the airlift. The productivity measurement involving the lowest flow rate and the smallest downhole pressure change was between 20.4 and 25.5 gpm/psi. The next largest flow rate change was at the end of the first airlift, and it produced a productivity value of 22.2 gpm/psi, the same as the average value of the cessation of artesian flow. The two largest flow rate changes at the start and stop of the second airlift give virtually identical and higher productivity values of 27.5 and 27.2 gpm/psi. Figure 44. Detailed flowing and static logs from PS-13-1 run in September 2014 with precision SMU logging equipment. The flowing log was run during artesian flows, and the depths were increased by 1.4 m to have exactly the same bottomhole depth as the static log, as this is the most important part of the hole for this discussion. 20 25 30 35 40 45 50 55 60 65 70 76.2 76.4 76.6 76.8 77 77.2 77.4 77.6 77.8 78 Depth (meters)Temperature C PS 13-1 Sept. 2014 Static and Flowing SMU Logs PS 13-1 9-7-14 Static SMU PS 13-1 9-18-14 Flowing SMUCemented CastingSolid CasingWell Screen85 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 52 These values are quite encouraging, as the well did not give lower productivity values as higher flow rate changes occurred, indicating that the well is capable of flowing at significantly higher rates. However, the values do not indicate that the temperatures seen during testing are sustainable over the long term. Table 4. Well productivity data Start Artesian Flow Stop First Airlift Start Second Airlift Stop Second Airlift Starting Flow Rate 0 172 65 300 Ending Flow Rate 60 – 75? 55 300 60? Change in Flow Rate 60 – 75? 117 235 240 Pressure Before Change 103.40 38.27 46.45 39.4 Pressure After Change 100.46 43.54 37.91 48.23 Change in Pressure 2.94 5.27 8.54 8.83 Productivity (gpm/psi) 20.4 – 25.5 22.2 27.5 27.2 The pressure record in PS-13-1 shows a 2 psi increase after 22:30 hours on September 16 (Figure 41). This increase reflects the thin cable holding the tool breaking, and the tool moving part of a meter downhole until it was held by the thicker aircraft cable, which turned out to be useful backup for the Kuster tool. 9.4 Temperature and Pressure Monitoring in PS-13-2 Two hours after the first airlift of well PS-13-1, a Kuster tool was hung in well PS-13-2 near a depth of 30 m to monitor its downhole temperature and pressure for a few days during the expected longer and more voluminous second airlift. This PS-13-2 record is exceptionally complex for a well that was not flowing (Figure 45). The start and stop of the second airlift is marked by sharp pressure changes of about 0.2 psi. No net longer-term pressure change occurred between the pressure prior to the airlift and pressures near the end of the monitoring period. During the airlift, a curious temperature increase and decline was recorded that requires a much deeper understanding of the hydrology to explain (Figure 45). Equally large or larger temperature changes occurred when the airlift was not in progress. 9.5 Temperature and Pressure Monitoring in PS-13-3 A Kuster tool was also hung in PS-13-3 after the first airlift of PS-13-1 to document the downhole pressure and temperature changes (Figure 46). This record shows a sharp 0.2 psi reaction to both the start and stop of the airlift. There is no longer-term net pressure change from the start of monitoring to the end. A tiny 0.05°C temperature rise was associated with the higher 86 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 53 flow that did not reverse after the airlift. Also, three tiny temperature spikes occurred after 20:00, close to one day apart, that are not understood. Figure 45. PS-13-2 pressure and temperature response during PS-13-1 flow testing. Figure 46. PS-13-3 pressure and temperature response during PS-13-1 flow testing. 70 70.5 71 71.5 72 72.5 73 42 42.2 42.4 42.6 42.8 43 43.2 43.4 43.6 43.8 44 9/15/2014 12:009/16/2014 0:009/16/2014 12:009/17/2014 0:009/17/2014 12:009/18/2014 0:009/18/2014 12:00Temperature CPressure (psig)Date and Time PS 13-2 Monitoring Sept. 15-18, 2014 During 300 gpm Airlift of PS 13-1 Pressure Start 300 gpm air lift Stop 300 gpm airlift Temperature Resume 60 gpm artesian flowFlowing 60 gpm artesian 78 78.1 78.2 78.3 78.4 78.5 78.6 78.7 78.8 78.9 79 51 51.1 51.2 51.3 51.4 51.5 51.6 51.7 51.8 51.9 52 9/15/2014 12:009/16/2014 0:009/16/2014 12:009/17/2014 0:009/17/2014 12:009/18/2014 0:009/18/2014 12:00Temperature CPressure (psig)Date and Time PS 13-3 Sept 15-18, 2014 Monitoring During 300 gpm Airlift of PS 13-1 Pressure Start 300 gpm Air lift Stop 300 gpm Air Lift Temperature Resume 60 gpm artesian flowFlowing 60 gpm artesian Flowing 300 gpm 87 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 54 9.6 Historic Hot Springs Temperature Monitoring A small Hobo brand temperature monitoring probe was placed in the discharge area in the historic hot spring pool at PHS during testing of the wells. The pool is located 750 feet northeast of well PS-13-1. During the testing period, the sensor was placed in the northwest corner of the pool, about 2 feet below the water surface (Figure 47). Researchers Chris Pike and Dick Benoit also used a presision temperature measuring probe owned by Southern Methodist University to measure temperatures in the bottom of the pool, inserting the probe several inches into the sandy bottom of the pool and recording the temperatures. A maximum temperature of 73°C (163°F) was encountered in the extreme eastern edge of the pool. The water temperature of the pool was monitored between September 9 and 18, 2014, with a brief interuption during the early morning hours of September 16 to download data. During the time that the temperature was being recorded, the hot spring pool was being used by the public for soaking and relaxation activities. Chris Pike, an ACEP staff member, monitored the temperature probe on a nightly basis to ensure that it was still in position. During a brief period Figure 47. The historic hot spring pool temperature was monitored during flow testing of PS- 13-1. 88 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 55 on September 12, the probe was removed from the spring. The data collected show that the pool temperature varied for unknown reasons, but mostly stayed between 38°C and 43°C (100°F– 110°F) (Figure 48). During the 300 gpm flow testing, the pool temperature dropped and did not stabilize and begin to rise again until after airlift pumping was stopped. During this time, the pool dropped to its coolest recorded temperature, below 34°C (94°F) (Figure 48). Further testing is needed to draw a difinitive correlation between the temperature of the hot spring pool and the flow of the wells. However, pumping of water from the shallow thermal aquifer likely impacts the flow of hot water into the pool. 9.7 Flow Testing Conclusions Well PS-13-1 was airlifted for over 7 hours at an average flow rate of 300 gpm, which represented about the largest flow that could have been achieved with available equipment. A Figure 48. Hot spring pool temperatures during the September 2014 flow testing. The lowest temperature recorded occurred when the greatest flow rates were being pumped from PS-13-1. 90 92 94 96 98 100 102 104 106 108 110 9/9/14 0:009/10/14 0:009/11/14 0:009/12/14 0:009/13/14 0:009/14/14 0:009/15/14 0:009/16/14 0:009/17/14 0:009/18/14 0:00Temperature (°F)Pilgrim Hot Springs Hot Pool Temperatures During Well Testing PS13-3 Flowing ~ 60 GPM Artesian PS13-1 Flowing ~60 GPM Artesian PS13-1 Flowing 300 GPM PS13-1 Flowing 172 GPM Erroneous Data Hourly Hot Pool Temeprature (°F) 89 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 56 longer flow test would have been more desireable, and help to better define the resource however, due to time and funding constraints it was not possible. We acknowledge this weakness and recommend future flow testing prior to substantial investment in anything other than small scale power generation. Repeated productivity measurements with flow rate changes of 60 to 240 gpm gave values of 20.4 to 27.5 gpm/psi which indicate a productive well. It is encouraging that the productivity values associated with the higher flow rates had the highest values. During the airlift, most of the fluid must have entered the wellbore in the main shallow thermal aquifer and flowed down through the sediments along the blank casing to enter the screened part of the well below 57.3 m. The airlift test impacted wells PS-13-2 and PS-13-3 nearby with a 0.2 psi pressure decline. Apparently, temperature impacts also occurred, but the indicators are not convincingly explicable with the available data. 10. PILGRIM GEOTHERMAL SYSTEM CONCEPTUAL MODEL 10.1 Conceptual Model History Conceptual speculation about the PGS dates to its first recorded visit by geologist G. A. Waring, who was interested in the geothermal system (Waring, 1917). Waring noted that the sulfate-to- chloride and the calcium-to-sodium ratios in the thermal water were much different from seawater and that the relatively high salinity was not due to “an admixture with sea water” (p.74). Also, Waring speculated that “beneath the river alluvium the bedrock may be gneiss, intruded by a granitic mass … the heated water rises along the fractured contact zone between the two kinds of rock”(p.75). Waring drew no schematic diagrams of the PGS. In the early 1970s, the USGS embarked on a program to improve the understanding of geothermal systems in the United States; it developed long-lasting conceptual understandings of the PGS even if it did not draw a specific conceptual model. In Alaska, Miller et al. (1975) assessed the geochemistry of many of the known springs and their regional geologic setting. In terms of regional setting, the proximal relationship between thermal springs and granitic plutons was recognized. Miller et al. (1975) presented the first stable isotope analysis of the Pilgrim thermal water, which showed the water to be derived from local rain and snowmelt. The thermal fluid is not a mixture of meteoric water and seawater. Miller et al. (1975) presented the first predicted PGS subsurface temperatures—137°C (279°F) based on the quartz geothermometer and 146°C (295°F) based on the Na-K-Ca geothermometer—and concluded that the thermal water must be at depths of 9000 to 15,000 feet (3.3 to 5.3 km) to reach these temperatures, based on a gradient of 30°C–50°C/km. Finally, Miller et al. (1975) observed that “most, if not all, of the hot springs are characterized by reservoirs of limited extent and relatively low temperatures in comparison with temperatures of geothermal systems presently being exploited for power generation”(p.12). The first conceptual model of the PGS was bravely put forth following the 1979 field season, with the focus on small-scale and very shallow convection cells and a water balance model (Osterkamp et al., 1980). The water balance model (Figure 14 from Osterkamp et al., 1980) allows for the possibility of rising hot water being impacted by subpermafrost cold recharge and 90 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 57 recirculation of thermal water. Various power estimates were made. The geothermometry from PHS and the first two wells drilled in late 1979 indicated the presence of a deeper and hotter 145°C–150°C (293°F–302°F) reservoir (Motyka et al., 1980). The second conceptual understanding of the PGS was developed in 1982 after the first six wells had been drilled and flow tested (Figure 3 from Economides, 1982). The primary result was the recognition of the shallow reservoir of laterally flowing thermal aquifer and the exceptionally bold conclusion was that “the existence of a hot water zone of about 150°C (302°F) and at a depth of around 5000 feet is now virtually certain” (Economides, 1982; Economides et al., 1982, p.30). In fact, a more specific depth of 4875 feet was stated, based on extrapolating the positive temperature gradients beneath the shallow thermal aquifer. The researchers also stated that “locating the hot water source for the shallow zone is relatively unimportant, since the fluid at depth provides a high temperature source formation extending aerially at least as far as the total area drilled” (p.28). Regrettably, the total area drilled by 1982 amounted to only a few square acres. Liss and Motyka (1994) relied upon geochemical data to suggest that Tertiary–Quaternary marine sediments might underlie PHS and that the PGS might have subsurface temperatures as hot as 190°C–230°C (374°F–446°F) based on a Mg-Li geothermometer and admittedly suspect noncondensible gas geothermometry. More recent drilling at PHS did not encounter Tertiary– Quaternary marine sediments. Liss and Motyka (1994) also noted a 3He/4He value of 0.9, which suggested a mantle component of helium. No work was performed on the PGS from 1993 until 2010. Following this nearly two-decade hiatus, Daanen et al. (2012) utilized the COMSOL Multiphysics finite element package to develop the first numerical model of the PGS. This modeling assumed steady-state conditions with an ongoing flow of cold water toward the geothermal system being required to maintain the high negative-temperature gradients beneath the shallow thermal aquifer. The model indicated that potentially 38 MW of thermal energy moves through the shallow groundwater system near PHS. Concurrently, Chittambakkam et al. (2013) utilized the TOUGH2 simulator and assumed similar steady-state conditions to estimate a total heat loss of 26 MW (Appendix N). Unfortunately, more recent geologic studies have brought new information to light which does not coincide with the results of these modeling efforts. In the case of Daanen et al. (2012), the observed vertical temperature distribution given in their Figure 3 shows no shallow lateral flow and implies that it should be possible to drill a well below PS-1 with near isothermal temperatures of 90°C (194°F), which was disproven by the drilling of well PS-13-3. In the case of Chittambakkam et al. (2013), their simulated temperatures in Figures 12 and 13 do not show 90°C water flowing from depth to the surface. Instead, there is an unexplained cooling and then reheating of the thermal upwelling. Benowitz et al. (2013) used thermochronology modeling to constrain the tectonic regime responsible for the PGS. They conclude that the thermal anomaly is related to the youthful extensional setting of the Kigluaik range front fault and is not thermally equilibrating, suggesting that the hottest temperatures have not been accessed (Appendix O). 91 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 58 Miller et al. (2013a) present the most complete and detailed conceptual picture of the upper 700 feet of the PGS (Figure 49). In their model, an areally restricted near-vertical thermal upwelling transmits 90°C water almost to the surface. There is a near-cylindrical thermal anomaly with a radius of 500 to 800 m extending outward from this upwelling that, by the distance of the temperature decline to 20°C (68°F), is basically vertical. Also shown in the Miller et al. (2013a) model (Figure 49) is a strong flow of cold groundwater beneath the shallow aquifer flowing toward the thermal upwelling from both east and west and then flowing north toward the Pilgrim River. This is the steady-state model wherein the thermal anomaly and static temperature profiles of the various wells could remain in the described condition for an indefinite period provided the relative flow rates of hot and cold water remain more or less constant. One other challenge for the model developed by Miller et al. (2013a) is that three recent wells drilled to locate thermal upwelling do not show it located where it is shown in the model (Benoit et al., 2014a). All subsurface temperature data acquired to date were used to create the plan view maps shown in Figure 8. These maps show temperature contours of the shallow thermal aquifer and the temperature minimum, measured from the deep wells. These data along with modifications to the Miller et al. (2013a) model were used to create Figure 51, which shows the current understanding of the PHS upwelling. The model in Figure 51 shows temperature contours across the thermal anomaly, using a northwest to southeast cross section. The upwelling is shown in the area northwest of PS-13-1, where no subsurface exploration has been attempted due to swampy conditions and challenging access. Numerous Geoprobe and temperature gradient holes south and east of PS-13-1 allow a Figure 49. Conceptual model from Miller et al. (2013a). 92 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 59 high degree of confidence in the subsurface temperatures. North and west of PS-13-1, no drilling activities have taken place, and the temperatures shown are estimates that could occur, based on the conditions given in the model. Glen et al. (2014) developed a more regional conceptual model of the overall geothermal system (Figure 50). This model consists of a diffuse downward flow of meteoric water through basement rocks, along range-bounding faults that separate the Kigluaik Mountains and Hen and Chickens Mountain from the Imuruk Basin. Hot fluid is shown diffusely rising through bedrock beneath the valley; it becomes focused in a narrow inferred northeast-trending structure that is diagonal across the basin, and then rises obliquely in a northeasterly direction. The proposed northeastward hot flow direction is largely based on a prominent gravity low southwest of PHS, which suggests 800 m depth to bedrock (Figure 22). Fluid then flows along the shallowing and narrowing bedrock contact toward the northeast, where close to the surface location of the hot springs it is further concentrated into north- and northeast-trending structures that allow it to rise steeply through approximately 300 m of clay-rich alluvium to the surface. This model is constrained by drill-hole data only in a small area near the hot springs. No temperature or quantitative depth distribution of the thermal fluid is shown in the Glen et al. conceptual model. Figure 50. Regional conceptual model cartoon from Glen et al. (2014). 93 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 60 The culmination of this report is to present as complete a conceptual model as possible of the PGS. This model uses to the maximum extent the previous efforts and covers the range from regional aspects to quite detailed local features within the thermal area. There is no doubt that the thermal fluid at PHS comes from a local meteoric source. Whether this fluid is recharged from the Kigluaik Mountains to the south or from the Pilgrim River Valley is unknown, as not enough local meteoric isotopic samples have been collected from these areas. Discriminating between these two possibilities will first require determining if there are measurable isotopic differences between precipitation falling on the south and north sides of the Kigluaik Mountains. Obviously, this information is of more academic than practical interest in trying to develop the PGS, which is why it was not pursued as an integral part of the recent exploration effort. To date, drilling efforts at PHS have been unsuccessful in finding the estimated subsurface temperatures of 140°C to 150°C (284°F–302°F) that have provided much of the impetus for extensive exploration of this location over the past 40 years. Perhaps the higher temperatures are a relatively large lateral distance away from the thermal springs or at much greater depths. If so, the direction to go is uncertain. Even if a location lateral to the thermal springs were known, the costs to access it by road would likely be high. Alternatively, perhaps the quartz and NA-K- 1/3Ca geothermometers may not have been appropriate for the PGS, and more conservative predictions, such as the chalcedony and Na-K-4/3Ca geothermometer, should have been used. In this case, the predicted subsurface temperatures would be near boiling at depth. Another possibility is that somehow the exceptionally high calcium content in the PHS water is impacting the accuracy of the cation geothermometers. A lower base temperature of perhaps 100°C (212°F) for the geothermal system does not require the meteoric water to descend to depths of 9000 to 15,000 feet, as calculated by Miller et al. (1975); it still would have to go as deep as 10,000 feet. Unfortunately, no background heat flow holes are anywhere near PHS to constrain the regional background temperature gradient. In west-central Alaska, Miller et al. (1975) observed that “apparently fracture systems were not developed or are not sufficiently open in well-foliated regionally metamorphosed rocks to allow deeply circulating hot water to gain access to the surface” (p.6). In the 40 years since this observation was first printed, very few producing geothermal fields have been hosted by foliated metamorphic rocks. The few examples of wells producing from foliated rocks are not highly productive. This makes it more challenging and risky to drill into the metamorphic bedrock beneath PHS to produce from fractured bedrock. To date, the bedrock samples recovered from PHS drilling have been metamorphic, not granitic. Granitic rocks and other volcanic and sedimentary rocks in west-central Alaska regularly host geothermal systems. However, no geophysical interpretation or discussion has yet argued granitic rocks are present beneath PHS. The structure(s) controlling the thermal fluid flow remain poorly understood, though there is agreement that the east–west-trending Kigluaik range-front fault is the dominant structure in the vicinity of the PGS, and is probably close to optimal orientation for the critical failure needed to create open space for thermal fluid flow. Structures trending north–south or northeast–southwest would be much less likely to pull apart to create the needed open space for fluid flow. Structures oriented in these less optimal directions would need to be in more complex local settings for 94 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 61 permeability. Intersections of single narrow faults present small targets and are unlikely to have enough areal extent for development as geothermal fields. This kind of geometry would require some second permeability, such as a connection to nearby permeable formations, to store enough fluid to develop a viable production/injection strategy. The dip of the Kigluaik range-front fault at depth is unknown, so it is highly speculative to draw it at a low enough angle for realistic penetration by a well near the thermal springs. If other east–west-trending faults are present beneath the Pilgrim Valley, they have not created a density contrast large enough to be recognized on the gravity map. It is interesting to speculate on the nature of the local structure or feature at PHS that is transmitting the thermal water from the top of the metamorphic rocks at 320 m to the surface. Nobody has yet described any surficial indication of such a structure, which must be either very young or somehow continuously active to maintain its permeability. Miller et al. (2013a) recognize a 1 m high north–south terrace that is near the west edge of the thermal anomaly near the postulated north–south-trending fault, but also note that the terrace could be the result of frost heaving. This feature must penetrate and keep open multiple layers of soft clay above the bedrock. It is difficult to describe a feature that has proven so elusive. The drilling and temperature results to date indicate this feature is most likely northwest of where drilling has occurred. If the zone of upwelling is located between the already drilled wells, then it is likely so small that whether it represents a viable target becomes a question. 10.2 Current Pilgrim Geothermal System Understanding With the background presented thus far, the conceptual model based on our current understanding of the PGS contains the following components: 1. Local meteoric water must travel to depths of 15,000 feet to provide a resource temperature of 150°C (302°F) if the regional temperature gradient is 30°C–50°C/km (Miller et al., 1975). If a lower resource temperature near 100°C (212°F) is present then the water may only need to travel as deep as 10,000 feet. Whether the cold water flows down a single fault in a concentrated manner or through myriad small fractures in bedrock is only of academic interest, as no developer’s activities are likely to impact this flow. 2. The area of the top hundred or so feet near the discharge point of the PGS has been quite well characterized. Some thermal water is actually able to reach the surface and discharge through the thermal springs. This water reaches the near surface with a temperature of 91°C (196°F). Some thermal water is discharged into a very shallow aquifer several feet or meters below the surface and spreads laterally over a fairly large area. Most of the thermal water is discharged through the shallow thermal aquifer near a depth of 100 feet. Most likely, this shallow aquifer is charged not too far to the northwest of well PS-13-1. Where the water that percolates through the shallow aquifer eventually travels is unknown, as this has not been the primary purpose of recent exploration. It is suspected, however that this water travels to local sloughs, natural hot springs, and the Pilgrim River. The amount of water discharged onto the surface and into the shallow subsurface amounts to about 20 to 40 MW thermal. 95 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 62 3. The nature of the feature or features allowing the thermal water to rise through the Quaternary alluvium is unknown. Individual faults or intersecting structures have been previously proposed, but as noted, the proposed ideas are somewhat questionable and nobody has yet strongly argued the case for these features. 4. The northeast thermal anomaly is real, but so little is known about it at this time that it is unclear if it is part of the same geothermal system as the PGS. If it is part of the same system, then the argument for a northeast-trending structure becomes stronger. If the northeast thermal anomaly is not actually part of the PGS, then the possible northeast trend might actually be misleading. 5. The sharply declining temperatures beneath the shallow thermal aquifer can be interpreted in two ways. A steady-state model requires that cold water be flowing beneath the shallow thermal aquifer to remove the heat. A transient model does not require such cold groundwater flow. Arguments against the steady-state model are based on two facts. First, the original static temperature profiles in the deeper wells were all very smooth and showed none of the complexity that moving water imparts to them. Once the wellbores connected various zones of permeability, then water movement in the static temperature profiles became obvious (Benoit et al., 2014b). Second, the hot and cold flowing water would be competing for the same permeability channels near the thermal upwelling, presumably in gravel and sandy layers. The thermal water clearly has enough pressure to flow up to the surface but there are no recognized cold springs near the thermal springs. If there were abundant cold water with artesian pressure coming from the Kigluaik Mountains, only a perfect seal or a near-perfect pressure balance could separate the two hydrologies. With the amount of gravel described by Miller et al. (2014a), this seems quite unlikely. An argument might be proposed that having 100 m of permafrost surrounding the thaw bulb offers the best available protection from invading cold water at shallow depths. To place the proceeding arguments into a picture, we know there must be a thermal upwelling, and by default, the most likely place for its location is a modest distance to the northwest of well PS-13-1 (Figure 51). The nature of the permeability in this channel is uncertain. Faults or fault intersections have been hypothesized, but no convincing evidence has been presented for verification. The thermal water must have circulated deeply within the metamorphic or metamorphic/granitic bedrock. It is not known at what depth the thermal water became concentrated into the focused flow we see near the surface. Whether it has diffusely flowed along the top of the bedrock or has risen as concentrated flow through the upper part of the bedrock is speculative. 96 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 63 Figure 51. The current conceptual model of Pilgrim Hot Springs as shown in a cross-sectional view looking from southwest to northeast. This model, which is based on all data acquired through September 2014, indicates that the main upwelling zone is in the swampy area northwest of PS-13-1. Bedrock is represented by the dashed horizontal line at approximately 320 m in depth. The dashed temperature contours represent areas where the temperatures have not been well measured. 97 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 64 11. EXPORTING GEOTHERMAL ENERGY TO NOME Like most rural areas of Alaska, Nome relies on a diesel microgrid for its electrical power. The diesel fuel that powers this grid is shipped long distances during a short period when the sea is ice free and is stored in bulk fuel storage tanks until it is used for home heating or electricity generation. Rural communities in Alaska face challenging logistics, limited infrastructure, and poor economies of scale. These factors coupled with high oil prices equate to expensive energy prices that make economic development challenging in many of Alaska’s rural communities. In 2008, the Nome Region Energy Assessment, funded by the U.S. Department of Energy and the National Energy Technology Laboratory, concluded that geothermal energy was a potentially economic option for the region, depending on the size of the power plant that the geothermal resource could support (Sheets et al., 2008). Following the success of the Chena Hot Springs project, a preliminary feasibility study was performed in conjunction with the Nome Regional Energy Assessment report (Dilley, 2007). To support landowners and the City of Nome in their decision-making process regarding possible development options, several studies related to the integration of 2 MW of geothermally generated electricity into the existing Nome grid and the economics of the project have been conducted in conjunction with the geothermal exploration described in this report. 11.1 Geothermal Power Economics In 2012, a private developer representing Potelco Power and Telecommunications expressed interest in developing a geothermal project at PHS. Potelco believed the project could be economically viable, and transmission infrastructure could be constructed between PHS and Nome if the resource could provide at least 2 MWe. Potelco created Pilgrim Geothermal LLC, under which development activities would take place. The City of Nome negotiated a power purchase agreement with Pilgrim Geothermal LLC to purchase 2 MW of geothermally generated electricity. To determine if this energy would be cheaper for their ratepayers than traditional electrical power generated with diesel generators, UAF economist Antony Scott modeled the price of Nome diesel versus the price of Arctic North Slope crude and then used this information to project possible future Nome diesel prices based on U.S. Department of Energy crude oil price predictions. The work provided a framework for decision makers as they weighed the pros and cons of integrating a geothermal generation source into the Nome grid. Indeed, from a utility point of view, the most compelling aspect of adding geothermal is the opportunity to reduce the price volatility that results from the fluctuating price of diesel (Scott, 2015). This was the first attempt to quantify diesel price risk in remote locations that receive only a few fuel deliveries per year. The full report is shown in Appendix Q. 11.2 Wind-Diesel-Geothermal Microgrid Modeling In 2013, Nome increased its nameplate wind power capacity to 2.7 MW to provide a portion of the annual average load of 4 MW. To model the effect of 2 MW of geothermal power, which had been negotiated as a take or pay power purchase agreement, UAF researchers created a time step simulation model using two years of Nome grid data. As with the economic analysis explained in 98 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 65 the previous section, the grid modeling was intended to serve as a guide to Nome decision makers as they decided if the integration of geothermally generated electricity was beneficial to the utility’s ratepayers. A non-load following 2 MWe geothermal generation scheme was modeled in conjunction with the installed wind and diesel capacity. The Nome utility wished to observe the effect of possible geothermal power on its ability to fully utilize the wind resource, which would soon be owned, maintained, and controlled by them, and minimize the wind that would need to be “dumped.” Researchers also modeled the utility’s ability to fully use the wind if it added smaller generators to the diesel powerhouse. According to VanderMeer and Mueller- Stoffels (2014), “adding to the diesel generator fleet to create smaller, more consistent differences between the combined capacities of diesel generator combinations resulted in less diverted wind energy, more displaced diesel generated energy, a higher diesel generator load factor, and more diesel generator switching” (p.4). This modeling allowed the utility to determine the value of adding geothermal generation, while still considering the decrease in performance due to increased switching and decreased load factor. 11.3 Transmission from Pilgrim Hot Springs to Nome The remote location of PHS, 60 miles north of Nome, complicates any future development of a power generation facility. While the site is accessible via road, the transmission infrastructure must be constructed from PHS to Nome if any power generated on-site is to be purchased and consumed in Nome. A transmission option that has been investigated in hopes of lowering the infrastructure cost of transmission in rural Alaska is a high-voltage direct current (HVDC) transmission line. Conventional alternating current (AC) transmission requires three- or four- wire transmission infrastructure, while HVDC transmission requires one or two wires. This could reduce cost through wire savings and reduced structural loads, requiring fewer poles and saving money in materials and construction time (Polar Consult Alaska, 2012). This research is discussed in detail in Appendix R. 12. LESSONS LEARNED The remote location of Pilgrim Hot Springs, short snow-free construction season, thick layer of Quaternary alluvial fill above the bedrock which included multiple permeable layers, and limited accessibility, necessitated careful operational planning. During the course of the research at Pilgrim Hot Springs it has become clear that some parts of the project were highly successful, and other elements of the project could have been done differently and more efficiently. Many of the lessons learned are detailed below in the hopes that future geothermal exploration in the area can benefit from this experience. These lessons learned include:  The combination of the aerial FLIR and optical remote sensing data used in conjunction with the geoprobe exploration was an efficient and relatively cost-effective way to define the extent and temperatures of the shallow thermal aquifer. A slightly larger geoprobe unit might have been able to consistently penetrate through the temperature maximum of the shallow thermal aquifer and enable data collection as deep as 200 feet without the need for a full drill rig. If this unit had been used in the springtime when temperatures were cool and the ground was still firm, exploration northwest of PS13-1 might have been possible. 99 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 66  FLIR surveys served the dual purpose of finding hot seeps and allowing the heat flux associated with the PGS to be modeled. Two surveys were performed, one in the spring and one in the fall time. The spring survey was the most useful for mapping areas of snowmelt that correspond to permafrost-free areas and anomalous vegetation growth not regularly found on the Seward Peninsula.  Flow testing wells at flow rates exceeding the natural artesian flows proved to be challenging at PHS given the remoteness of the location, cost limitations, and temperatures. Basic water well pumps were not rated for the temperatures encountered at Pilgrim and required a rig on site to install. The lightweight air lifting apparatus that was utilized worked very well, however it did require close monitoring and regular refueling every couple hours. Flowing the well at rates greater than 300 gpm would have required additional manpower and equipment. The 6-inch Krohne magnetic flow meter that was used during this flow test was extremely reliable, effective, and accurate as well as user friendly.  The 525 ppm of calcium in the Pilgrim water is an obvious suspect in raising the question of whether thermal waters with exceptionally high calcium contents provide accurate geothermometry calculations. Given the inability to encounter the temperatures predicted by established geothermometry techniques, additional research is warranted to see if there are other geothermal systems where high levels of calcium caused overly optimistic geothermometry estimates.  The drilling associated costs consumed the most time and financial resources. The legacy wellhead repairs allowed researchers to utilize the wells that were drilled in the 1970’s and 1980’s for temperature and flow testing, and improved our understanding of the field. In addition, we were able to characterize subtle long term changes that have occurred over time. The temperature gradient slim holes were an economic way to measure temperatures at the top of bedrock. Ideally one or two more additional slim holes would have been drilled near where PS13-1 was eventually drilled and northwest of this area to help guide the later drilling of the large diameter well capable of flowing large quantities of geothermal fluid.  The Alaska Oil and Gas Conservation Commission regulates geothermal exploration and drilling in the same way as oil and gas associated activities are regulated, regardless of depth, or the temperatures and pressures that one is likely to encounter. Applying the same standards that the oil and gas industry is bound to for low temperature geothermal exploration creates a situation where this type of exploration becomes cost prohibitive and disincentivizes the development of small geothermal resources. Revisiting the regulations associated with geothermal exploration in Alaska is warranted. 13. CONCLUSIONS Geothermal exploration at Pilgrim Hot Springs (PHS) has significantly increased the understanding of this resource and enabled the planning of possible next steps, which are being carried out at the time this report is being written.  Initial project planning included the consideration of helicopter supported drilling based on the belief that the area of upwelling could be north of the Pilgrim River. Slim hole drilling was able to define the edges of the shallow thermal aquifer and the depth of the 100 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal Geothermal Exploration of Pilgrim Hot Springs, 2010-2014 Final Report 67 deep aquifer at PHS. Bedrock was encountered on several occasions which had not occurred during previous drilling at the site.  The possible upwelling area was constrained by drilling and the most likely upwelling zone feeding PHS is located slightly northwest of well PS13-1 (Figure 51). This idea is supported by the plan view temperature maps in Figure 8, which describe the flow direction of the thermal fluids.  We continue to believe the site is capable of supporting two megawatts of electrical power generation. The economic viability of exporting this power to Nome remains a question that private industry is best suited to answer. Repeated productivity measurements of well PS13-1with flow rate changes of 60 to 240 gpm gave values of 20.4 to 27.5 gpm/psi which indicate good productivity. A longer flow test would have been more desireable, and helped to better define the resource, however, due to time and funding constraints it was not possible. We acknowledge this weakness and recommend future flow testing prior to substantial investment in anything other than small scale power generation.  The landowners at PHS are investigating different on-site development options and are currently moving forward with plans to begin producing agricultural products on the site. This could include the construction of greenhouses to produce food for export to local communities, tourism infrastructure, and community facilities. Traditionally, the export of geothermal power has occurred in the form of electricity; however, using the heat energy at PHS to grow food for the region could be a creative way to export the “energy” and supply a much-needed commodity that otherwise is shipped into the region. The high cost of food transported to the area is heavily impacted by the price of petroleum.  Pilgrim Geothermal, LLC has indicated that it is planning additional drilling activities to identify the future production-well location for a large-scale geothermal electric power plant. Future exploration could rely on angle drilling from the existing drill to access the northwest target area. 101 of 101 Kawerak-Pilgrim Hot Springs-REF Proposal