Drilling For Superhot - Geothermal Energy
Drilling For Superhot - Geothermal Energy
Drilling For Superhot - Geothermal Energy
Dr. Rebecca Pearce is a Fellow on the Energy Systems team and the Ultradeep Geothermal Science Lead at the
Cascade Institute. Her research focuses on geothermal site characterization and emerging geothermal and drilling
technologies. Rebecca is an expert in applied magnetotellurics and has contributed to projects in Northern British
Columbia, Chile, and West Antarctica. She holds a PhD in Geophysics from University College London.
Tony Pink is the Owner/Director of Pink Granite Consulting LLC and was recently appointed the Task Force
Chairman for Drilling Technology for the US Department of Energy GEODE project. He has 34 years of drilling
experience in oil and gas and geothermal. Before starting his own business, he was VP of Subsurface Technology
for NOV focusing on the development of tools and equipment for the geothermal industry. Tony has a BSc in
Geology from UCW Aberystwyth and a postgraduate diploma from Harvard Business School. He has five patents
(two granted and three pending).
Acknowledgements:
We would like to thank Chris Rush for key contributions to the literature review and writing process; Jenna Hill,
Terra Rogers, Jordan Bendall, and Scott Janzwood for reviewing and editing the report; and our network of
experts from the international geothermal and drilling communities for reviewing the report and providing
helpful feedback.
Citation:
Pearce, R. and T. Pink. 2024. Drilling for superhot geothermal energy: A technology gap analysis. Version 1.0. Clean
Air Task Force and Cascade Institute. https://cascadeinstitute.org/technical-paper/drillingreport/.
The views expressed herein are those of the authors and do not necessarily reflect the views of the Clean Air
Task Force or Cascade Institute. This report has been simultaneously published by the Clean Air Task Force under
the title Bridging the Gaps: A survey of methods, challenges, and pathways forward for superhot rock drilling,
which can be accessed at: https://www.catf.us/superhot-rock/bridging-gaps/
1
Summary
The research frontier of drilling and well construction for superhot rock (SHR) geothermal energy systems—the
production of renewable, baseload electricity by circulating water in deep (>5 km), hot (>374˚C) rock—is steadily
advancing. Recent achievements in polycrystalline diamond carbide (PDC) drill bit design, improved rates of
penetration (ROP) into hard rock, and the development of insulated drill pipe show that deep drilling for SHR
geothermal projects is on the not-distant horizon.
But several key technology gaps still stand in the way of deep drilling in hostile subsurface geological
environments. Technology companies and laboratories must make rapid advances in specialized drilling rigs, bit
technology, high-temperature downhole tools, and temperature management equipment. Currently, these
drilling systems—and the amount of time required to access deep, hard rock formations—create significant
project costs. To bring SHR geothermal to commercial viability, technology companies and laboratories must
rapidly develop, test, and deploy new technologies.
This report reviews state-of-the-art deep geothermal drilling and well-construction technologies, identifies
existing technology gaps, and suggests strategies to overcome these gaps. Each technology is given a technology
readiness level (TRL) between 1-9, from theoretical to commercially scalable.
Overall, we find that SHR geothermal wells can be drilled by deploying a combination of existing technologies and
that the technological challenges to SHR drilling are surmountable. The economic challenges are a function of
limited availability and testing of these drilling systems, which will decrease as the SHR geothermal industry
expands.
A first-order gap shared by these technologies is the lack of access to SHR conditions, both in-field and in controlled
laboratory conditions. Without open-access experimental facilities and pilot sites, these technologies cannot
undergo iterative improvements necessary to de-risk SHR drilling and propel the industry forward.
Table of Contents
Summary.....................................................................................................................................................................2
Acronym glossary .......................................................................................................................................................6
1. Introduction ........................................................................................................................................................7
2. Background .......................................................................................................................................................10
2.1 Drilling challenges...........................................................................................................................................10
2.2 Frontier pilot sites ..........................................................................................................................................11
Utah’s Frontier for Research in Geothermal Energy (FORGE) ..........................................................................11
Fervo Energy (Blue Mountain, Nevada; Cape Station, Utah) ...........................................................................12
Iceland Deep-Drilling Project (Krafla and Reykjanes, Iceland) .........................................................................13
St1 Deep Heat Project (Otaniemi, Finland) ......................................................................................................14
Conclusions .......................................................................................................................................................15
3. Drilling system requirements for SHR ..............................................................................................................17
3.1 Requirements for rock-destroying equipment ...............................................................................................17
3.2 Requirements for high-temperature downhole tools ....................................................................................17
3.3 Requirements for temperature management equipment .............................................................................17
3.4 Requirements for casing, cement, and corrosion inhibition ..........................................................................18
3.5 Requirements for topside equipment ............................................................................................................19
4. Rock-destroying equipment: Technology frontier and gap analysis ................................................................20
4.1 Conventional rotary drilling ............................................................................................................................20
4.1.1 ROP generation........................................................................................................................................21
4.1.2 High temperatures...................................................................................................................................24
4.1.3 Directional drilling ...................................................................................................................................26
4.1.4 Performance in interbedded formations.................................................................................................27
4.1.5 Gaps and challenges for SHR ...................................................................................................................29
4.2 Hybrid conventional drilling methods ............................................................................................................30
4.2.1 Percussive drilling ....................................................................................................................................30
4.2.2 Waterjet drilling.......................................................................................................................................34
4.2.3 Particle drilling .........................................................................................................................................35
4.3 Direct energy methods ...................................................................................................................................38
4.3.1 Plasma drilling .........................................................................................................................................38
3
4.3.2 Millimeter wave drilling ...........................................................................................................................41
4.4 Conclusions for rock-destroying equipment ..................................................................................................45
5. High-temperature downhole tools: Technology frontier and gap analysis .....................................................48
5.1 High-temperature tool manufacturers ...........................................................................................................49
5.2 High-temperature motors ..............................................................................................................................53
5.3 Conclusions .....................................................................................................................................................53
6. Temperature management equipment: Technology frontier and gap analysis ..............................................54
6.1 Insulated drill pipe ..........................................................................................................................................54
Titanium drill pipe.............................................................................................................................................57
6.2 Drilling fluid ....................................................................................................................................................59
6.2.1 Water-based fluid ....................................................................................................................................59
6.2.2 CO2-based fluid ........................................................................................................................................61
6.2.3 Optimizing fluid dynamics .......................................................................................................................62
6.2.4 Low heat coefficient coatings ..................................................................................................................62
6.2.5 Mud Coolers ............................................................................................................................................64
6.3 Conclusions .....................................................................................................................................................65
7. Corrosion inhibition: Technology frontier and gap analysis.............................................................................66
8. Conclusion ............................................................................................................................................................68
Sources .....................................................................................................................................................................69
Appendix ...................................................................................................................................................................74
Technology specific findings: ................................................................................................................................74
Technology gap summary table ...........................................................................................................................77
4
Table of Figures
Figure 1 Specific enthalpy of pure water under increasing temperature and regimes .............................. 7
Figure 2 Depth versus on-bottom drilling hours for the wells drilled at Utah FORGE as of 2023 ............ 12
Figure 3 Total depth drilling times for the 8 wells at the Cape Station site in Utah ................................. 13
Figure 4 Geological diagram of the Otaniemi EGS site in Finland, including well location, depth, and
120C̊ temperature horizon. ....................................................................................................................... 15
Figure 5 Polycrystalline Diamond, Roller Cone and Hybrid PDC + Roller Cone bits. From the Baker
Hughes VulcanixTM Geothermal bit line .................................................................................................... 21
Figure 6 WOB and RPM regimes conditions for stick-slip dysfunction for conventional versus REVitⓇZT
drilling methods. ....................................................................................................................................... 24
Figure 7 Nozzle configuration to optimize drill bit cooling with fluid flow ............................................... 25
Figure 8 Profile of curved section of well 16B, showing correlations of hole smoothness and rotary-
motor systems used .................................................................................................................................. 27
Figure 9 Carbon Diamond Element Drill Bits ............................................................................................ 28
Figure 10 Schematic of the HydroVolve GeoVolve HYPERDRIVE system .................................................. 32
Figure 11 Orchyd and Drillstar hydro-mechanical drill system, and the bit-jet-rock interaction at the
drilling interface. ....................................................................................................................................... 33
Figure 12 Breakdown image of a radial jet drilling system within a borehole. ......................................... 35
Figure 13 Latest deign of 8 ½" particle impact/PDC hybrid bit. ................................................................ 36
Figure 14 A picture of drilling rig VERMEER 55 in the tunnel at VersuchsStollen Hagerbach (VSH) ........ 37
Figure 15 Visual representation of the two different versions of plasma-based fragmentation ............. 39
Figure 16 Schematic and real demonstration of plasma drill bit drilling through hard rock .................... 39
Figure 17 Simplified diagram of gyrotron-powered MMW drilling mechanism ....................................... 42
Figure 18 Ablation front, vitrified granite, stand-off distance, copper-steel waveguide .......................... 42
Figure 19 Simulation results for multi-strata with heterogeneous absorptive properties, showing that
the vaporization process does not produce a uniformly radial well......................................................... 44
Figure 20 Picture of Hephae’s high-temperature integrated MWD and RSS. ........................................... 49
Figure 21 A conventional positive displacement drilling system. ............................................................. 51
Figure 22 Temperature profiles when applying IDPs along (Left) the upper half of the wellbore and
(Right) the lower half of the wellbore with drilling mud. ......................................................................... 56
Figure 23 IDP design from NOV that allows an inner diameter of 3.69” inside a 5 7/8” drill pipe tube. . 57
Figure 24 Images of titanium drill pipe from ALTISS Technologies. .......................................................... 58
Figure 25 Drilling fluid shear strength [lbf/100ft2], plastic viscosity (PV) [cp], yield point (YP) [lbf/100ft2]
after exposure to 56C̊ and 300C̊................................................................................................................ 60
Figure 26 Low heat coefficient coating for drill pipe. ............................................................................... 63
Figure 27 Thermal conductivity comparison between TK34XT and the low heat coefficient coating TK™-
340TC/18TC. .............................................................................................................................................. 63
Figure 28 Impact of mud surface temperature by a mud cooler on BHCT of wells A & B. ....................... 65
Figure 29 NOV’s Glass Reinforced Epoxy corrosion-resistant casing liner. ............................................... 67
5
Acronym glossary
Adjustable kick-off (AKO) Mechanical specific energy (MSE)
Advanced Research Projects Agency-Energy (ARPA-E) Measurement-while-drilling (MWD)
Bottom hole assembly (BHA) Metal-to-metal (M2M)
Bottom hole circulation temperature (BHCT) Millimetre wave (MMW)
Closed-loop geothermal (CLG) Nth-of-its-kind (NOAK)
Conical diamond element (CDE) Non-productive time (NPT)
Conventional drill pipe (CDP) Oak Ridge National Laboratory (ORNL)
Depth-of-cut (DOC) Oil & gas (O&G)
Department of Energy (DOE) Particle Drilling (PD)
Directional steel shot drilling (DSSD) Polycrystalline diamond carbide (PDC)
Drilling in Deep Super-Critical Ambient of Positive displacement motor (PDM)
Continental Europe (DESCRAMBLE) Rate of penetration (ROP)
Enhanced geothermal system (EGS) Rotary steerable system (RSS)
First-of-its-kind (FOAK) Superhot rock (SHR)
Finite element analysis (FEA) Supercritical carbon dioxide (sCO2)
Geothermal Anywhere (GA) Technology readiness level (TRL)
German Continental Deep Drilling Programme (KTB) Torque on bit (TOB)
High electron mobility transistors (HEMTs) Total depth (TD)
High-temperature, High-pressure (HTHP) True vertical depths (TVD)
Inner diameter (ID) Utah Frontier Observatory for Research in
Insulated drill pipe (IDP) Geothermal Energy (FORGE)
Integrated ceramic electronics (ICE) Versuchsstollen Hagerbach (VSH)
Iceland Deep Drilling Project (IDDP) Water-based mud (WBM)
Levelized cost of electricity (LCOE) Weight on bit (WOB)
Log-while-drilling (LWD)
6
1. Introduction
The expansion of geothermal heat use for energy production beyond its current tectonic boundaries offers an
opportunity to supply carbon-free, renewable, baseload power anywhere in the world. It is estimated that two
percent of thermal energy within 3-10 km of the surface may offer the equivalent of 2,000 times the current
energy demand in the US and is virtually inexhaustible (Petty et al., 2020; Tester et al., 2006; van Oort et al., 2021).
This vast resource of heat can be extracted through the circulation of water in hot, naturally or artificially
fractured, low-permeability rock formations referred to as an Enhanced Geothermal System (EGS), or from a
subsurface interconnected closed-loop well network, referred to as Closed-Loop Geothermal (CLG). However,
these “next-generation geothermal systems” are not cost-competitive with mature energy technologies. Drilling
is the most cost-intensive process of a geothermal project and current drilling systems require further innovation
to perform in the hostile geological domains targeted by these projects. Advancing drilling technologies is critical
to make next-generation geothermal projects economically and technically feasible.
The ideal application of next-generation geothermal systems is the production of electricity with high-enthalpy
(>2,100 kJ/kg) water under supercritical conditions (specifically, >374˚C temperatures and >22 MPa pressures in
pure water—Figure 1). Researchers estimate that a levelized cost of electricity (LCOE) of USD 25-45/MWhr can be
achieved with an EGS flow rate of 80 L/s from high-enthalpy water, sourced from SHR (Cladouhos & Callahan,
2023). At these conditions, a geothermal plant with a 1 km 2 footprint and a tri-well system consisting of one
injector and two producers may sustain an 80-100 MW capacity for the duration of its lifetime (40-80 years) (Petty
et al., 2020). This disruptive technology is finally receiving recognition and, within the past few years alone,
research programs have rapidly advanced the R&D frontier of SHR technologies.
Figure 1 Specific enthalpy of pure water under increasing temperature and regimes
The fluid state and wellhead temperatures required for low-temperature conventional and SHR geothermal plays (Cladouhos & Callahan,
2023)
7
One of the biggest technical obstacles to achieving scalable, cost-competitive SHR is well construction and
completion in lithologies at superhot conditions. The average geothermal gradient in the world is 30˚C/km but can
range from 15-20˚C/km in shield-type cratons and subduction zones to 40-100˚C/km in volcanic regions or
extensional tectonic regimes. Assuming an average global surface temperature of 15˚C, the typical depth to
supercritical temperatures is in the range of 7-10 km beneath the surface into crystalline basement (van Oort et
al., 2021). However, the depth range accounting for the global variability in geothermal gradient is 5-20 km.
To successfully construct a geothermal well, an SHR drilling program must overcome three overarching
challenges:
Overcoming these challenges will require highly specialized drilling, casing, cementing, and completion
equipment, some of which exist today and others that require further research and financial investment (Petty et
al., 2020; Thorsteinsson et al., 2008). Industry, along with university and government research labs, are building
upon the decades of research conducted by the oil and gas (O&G) industry on the construction of complex, high-
temperature, high-pressure wells. Hydraulic fracturing at >5 km depth is not uncommon in the hydrocarbon
industry in sedimentary systems, and wells with over 15 km measured depth have been achieved. Wellhead
temperatures over 350˚C can be encountered in O&G and some high-temperature well completion equipment
exists (van Oort et al., 2021). While these technologies are transferrable to SHR, there are several key differences
between an ultradeep SHR geothermal well and a hydrocarbon well:
• SHR wells will reach >374˚C temperatures, whereas hydrocarbon wells typically range from 100-220˚C,
• SHR wells will be constructed in hard, abrasive, impermeable, crystalline rock with compressive
strengths of >240 MPa-far greater than soft, sedimentary rocks of <80 MPa normally encountered in
hydrocarbon wells,
• SHR wells will require larger production hole sizes to support higher flow rates,
• SHR fracturing may lead to severe losses of drilling fluid and may require wellbore sealing methods in
harder rock and higher-temperature conditions than experienced in hydrocarbon wells,
• Drilling and well completion programs for a SHR geothermal project typically account for 40-60 percent
of total project costs, and
• Very few geothermal wells are drilled annually relative to hydrocarbon wells, offering less opportunity
for iterative design improvements (van Oort et al., 2021) and the performance improvements seen in
shale drilling in the U.S.
8
These key differences, along with the challenges of well construction in supercritical conditions, create four
primary technology gaps that must be overcome to make SHR technically and economically viable—the first two
of which pertain to drilling and well construction:
To date, the drilling and completion programs for next-generation geothermal projects can account for 40-70
percent of total project capital expenditure costs (van Oort et al., 2021, Song et al. 2023). These costs may
substantially increase as deeper, hotter, and higher-pressure regimes are explored. While it is expected that the
first-of-a-kind (FOAK) demonstration sites will incur high upfront costs required to de-risk SHR through
experimentation, Nth-of-a-kind (NOAK) plants may achieve a cost-competitive LCOE as a result of the high energy
per well delivered to the surface by superhot steam and supercritical fluids, unleashing this vast reserve of energy
for sustainable power production.
This report investigates state-of-the-art drilling technologies striving for SHR temperatures by drilling ultradeep
(>5 km) wells through hard rock lithologies. The sections cover rock-destroying technologies, high-temperature
electronic tools, and temperature management equipment. Each review defines the TRL status of each technology
within each category, the technology gaps between their current capacity and the needs of SHR, and strategies to
overcome these gaps. The analysis in this report was conducted through an extensive literature review, dozens of
interviews with experts across the industry, and decades-long experience by the lead author of this study.
Section 2 begins by closely examining the overarching challenges for SHR identified above and describes the
landscape of next-generation experimental and pilot sites laying the foundation for further SHR research. Section
3 defines the requirements of an SHR well, including the drilling system, MWD and downhole sensors, top-side
equipment, drill string, and mud cooler requirements. Sections 4-7 review state-of-the-art rock-destroying
equipment (including subsets of conventional, hybrid, and novel direct energy drilling methods), high-
temperature electronic tools, downhole temperature management equipment, and corrosion inhibition
technology. These sections identify key gaps for SHR and strategies to address them.
9
2. Background
2.1 Drilling challenges
To reach SHR, a geothermal system must contend with hard crystalline basement rock, unprecedented depths
(from 5-15 km below the surface), and temperatures of >374˚C superseding the pressure-temperature capabilities
of most conventional drill bits (Naganawa 2017). In the case of conventional drilling, the casing and cementation
process is a major challenge in these conditions due to the extreme depths that require a larger hole diameter,
and the risk of casing/cement failure due to high temperatures and corrosive substances downhole, which could
potentially lead to partial borehole collapse, steam and brine leakage, and severe damage to the installed
equipment.
To accommodate these conditions, specialized borehole designs and temperature-rated cement and casing
materials are required to maintain well integrity, leading to significant costs (Petty et al., 2020; Thorsteinsson et
al., 2008). Downhole electronics, such as directional tools and measurement-while-drilling (MWD) tools, are rarely
rated above 200˚C (see Section 5). Lastly, in SHR conditions of >374˚C, rock may exhibit ductile behaviour and will
not fail under the rock-crushing mechanisms used by conventional drilling (Naganawa 2017). Corrosion is probably
less of a concern in SHR next-generation systems than hydrothermal systems, but the elevated temperatures will
still have implications for corrosion inhibition.
Several EGS and conventional geothermal projects have experienced complications due to the limitations of
conventional drilling and well-construction methods. While these are not SHR projects, they experienced similar
challenges to those stated above. Compromised well integrity has led to failed stimulation tests at Reykjanes
(Iceland), Newberry (Oregon), Sumikawa (Japan) and Beowawe (Nevada) sites (in 2017, 2012, 1989, and 1983
respectively), where stimulation fluids leaked through damaged or perforated casing. Fluid leakage can occur due
to thermal contraction from heating and cooling cycles during stimulation, or from cave-ins in deviated or
horizontal sections.
High pressures have led to the loss of equipment downhole, such as in the Cooper Basin (Australia) in 2009, where
a drill string was stuck at 3,700 m depth due to the pressure differentials between an over-pressured fracture and
the wellbore. The continued accumulation of mud and cuttings further prevented its recovery and the project was
terminated. Similar “differential-sticking” incidents occurred at The Geysers (California), Soultz-sous-Forêts
(France), Basel (Switzerland), Bad Urach (Germany), Soda Lake (Nevada), and Coso (California) geothermal sites,
some of which were ultimately abandoned due to this issue (Pollack, Horne, and Mukerji, 2020).
Over-pressure can also lead to kicks (sudden increases in well pressure) and blowouts, which were repeatedly
experienced in the Cooper Basin (Ibid.). Conversely, low-pressure, permeable formations that can be found in
geothermal domains, such as fractured rhyolites, granites, or volcanic tuffs, commonly lead to circulation loss of
high-pressure drilling fluid. Circulation loss can be mitigated by cement plugs and fluid influx control; however,
these strategies can incur costly drilling delays, even up to 10 percent of the drilling program costs (Denninger et
al., 2015). At the Utah Frontier Observatory of Research for Geothermal Energy (FORGE), Fenton Hill (New
10
Mexico), Bad Urach, and Soultz-sous-Forêts, zonal isolation equipment (specifically, packers) was prone to rupture
in high-temperature environments, which led to leakage and compromised well stimulation (Pollack, Horne, and
Mukerji, 2020).
Although these same challenges were mostly encountered at EGS and conventional geothermal projects,1 these
complications are also expected for SHR drilling. Several SHR boreholes have been drilled to date; however, no
SHR well has reached commercial operability. Historical SHR projects, such as the Drilling in Deep Super-Critical
Ambient of Continental Europe (DESCRAMBLE) project (Italy), are groundbreaking examples of the potential of
mechanical drilling technologies to access SHR conditions. However, the drilling technologies used in historical
SHR sites are no longer up to date. Therefore, a review of current EGS pilot sites provides a more complete
understanding of the technology frontier of SHR drilling.
Several recently completed and ongoing EGS pilot sites have made significant strides to advance the research
frontier of well construction in hot, hard rock (some in SHR conditions). Projects with relevant learnings include:
the US Department of Energy (DOE)-funded Utah FORGE, the commercial venture Fervo Energy(USA), the Iceland
Deep Drilling Project, and the St1 Deep Heat Project (Finland). Although there are other important sites, these
four projects were reviewed because of their timeline (recent drilling technologies were deployed), as well as their
relevant learnings for drilling in SHR.2
The wells drilled at the DOE-funded Utah FORGE project have been instrumental in improving conventional drilling
technologies for drilling into high-temperature impermeable lithologies up to 240˚C. The objective for Utah FORGE
is to push the frontier on hard rock drilling and reservoir creation for next-generation geothermal projects. While
not specifically pursuing SHR conditions, their achievements in ROP generation are directly transferrable to SHR
projects, as efficient drilling is essential to improving the economic feasibility of SHR. To date, 6 wells (Figure 2)
have been drilled, each demonstrating iterative improvements in ROP and bit-life that demonstrate the potential
of conventional drilling methods for SHR projects.
1 Please refer to Reinsch et al. (2017) for a thorough review of SHR drilling challenges encountered at Lardarello (Italy), Krafla (Iceland),
Kakkonda (Japan), the Geysers (USA), Los Humeros (Mexico) and Menengai (Kenya).
2 Please refer to Bromley et al., 2020 (and the references therein) for a description of DEEPEGS, Kakkonda, GEMex, and The Geysers, that
11
Figure 2 Depth versus on-bottom drilling hours for the wells drilled at Utah FORGE as of 2023
(Noynaert, GRC 2023)
Based on interviews with the project’s managing principal investigators, tests at FORGE revealed that drilling into
hotter, higher-pressure regimes such as SHR would require better systems for downhole management, including
the management of Mechanical Specific Energy (MSE) and vibrational dysfunction, better downhole sensing and
drilling automation, and improved rotary steerable systems. SHR drilling needs further advances in bit design and
MSE management to overcome interfacial severity from heterogeneous, interbedded formations, such as acoustic
and sonic techniques used in hydrocarbon exploration. Vibrational dysfunction is also a concern that requires
further innovation of topside equipment. “Look-ahead” technologies that locate potential blowouts (such as from
severely fractured lithologies) would allow the adjustment of parameters before hitting fractured domains.
Automated drilling may be a solution for bit dysfunction management. For example, FORGE saw performance
improvements with the Nabors REVitⓇ ZT automation program for downhole torque management. Meanwhile,
technologies like Thrubit™ logging, vibrational management, and ultrasonic imaging could be used for fracture
mapping. Rotary steerable systems for directional drilling have great potential but must be adapted for the harsh
environment of SHR. Next-generation drilling should also be applied in other geological settings, including hotter,
impermeable domains such as the Salton Sea, Coso, and the eastern US.
Almost all the equipment and methods used at FORGE can be used to drill over 300˚C, but the main limitation is
the temperature ratings of downhole sensing equipment.
Fervo Energy is not yet in pursuit of SHR; however, their learnings around ROP generation and the challenges from
their drilling programs are highly relevant for future SHR projects. In 2023, Fervo achieved major improvements
in ROP at its site in Cape Station, Utah. At the time of publication, Fervo has drilled two horizontal wells in Nevada
and six in Utah. The wells were drilled from 2022-2024 to approximately 4 km measured depth, with 900 m lateral
sections (Figure 3).
12
Figure 3 Total depth drilling times for the 8 wells at the Cape Station site in Utah
(El-Sadi et al., 2024)
Drilling fluid temperatures were monitored throughout the drilling operation in mud pits and through MWD. The
maximum temperature downhole was 107˚C, significantly lower than the static reservoir temperature of 109-
210˚C, with no MWD or directional tool failures (Fercho et al., 2024). In the context of SHR, these results indicate
that cooling the borehole while drilling is a worthwhile approach for MWD and LWD procedures and should be
adopted for future projects in hotter regimes. In an interview with Fervo engineers, they identified three main
gaps for SHR:
• Friction in the lateral wells may lead to accelerated bit wear and energy losses;
• While drilling to <450˚C is not a major concern, downhole chemistry (see Section 7) and high-temperature
drilling tools require further innovation; and
• ROP generation will likely reach a limit once wells can be drilled in 10 days, at which point drilling is not a
major cost factor (however, Fervo’s wells are shallow and cool relative to SHR wells).
The Iceland Deep-Drilling Project (IDDP) is a component of the DEEPEGS initiative and is funded by HS Orka,
Landsvirkjun, Reykjavik Energy, the National Energy Authority of Iceland, Alcoa, Equinor, and the EU Horizon 2020
program. Their first well (drilled in 2009 at Krafla, IDDP-1) was intended to be 4,500 m but encountered molten
lava at 2,100 m, which produced steam at 452˚C and made it the hottest production well on record. Completed in
2015, it was drilled with 177 - 288˚C temperature-rated tricone bits, a conventional motor, and high-temperature
metal-face seals (Friðleifsson et al., 2019). The well was ultimately decommissioned due to high levels of corrosion
and abrasion of the downhole equipment (Petty et al., 2020). IDDP-1 could have been the world’s first magma-
13
sourced EGS system, however, persistent valve failures resulted in the closure of the well. Further innovation of
wellhead systems to cope with high pressures are needed for SHR.
The second well (IDDP-2) was drilled at Reykjanes in 2016 by extending an existing conventional well from 2,500
m to 4,659 m (MD) with a 30o tangent starting at 2,750 m and reaching bottom hole temperatures of 426˚C and
340 bar pressure—truly supercritical for both fresh and saline water. There were substantial circulation losses
below 3.2 km to the final depth and the MWD temperature sensor was placed 14 m above the bit (Friðleifsson et
al., 2019; Stefánsson et al., 2018). The estimated true temperature was closer to 540˚C (Petty et al. 2020).
The demonstration well IDDP-2 proved an exceptional resource of geothermal energy, increasing the productivity
of the existing site by 40-50 percent. IDDP primarily used high-temperature roller cones and one hybrid PDC/roller
cone (further discussed in Section 4.1). The geothermal field was artificially stimulated at much greater depths
and temperatures than the existing conventional well using specialized drill bits and BHA technologies (Friðleifsson
et al., 2019). This first-of-its-kind geothermal well is a major achievement for the development of supercritical
geothermal resources. Runs 12 and 13 included a pioneering metal-on-metal motor designed and built by Baker
Hughes.
While IDDP-2 was a groundbreaking accomplishment in SHR drilling, circulation losses proved to be a serious
complication. Pumping water through the drill string is essential to maintain manageable bottom-hole
temperatures, requiring 40-60 L/s. Water must also be pumped into the annulus to cool the casing. Losses started
at 3.2 km to the final depth, after which point new water resources had to be continuously pumped, resulting in
the consumption of 1.5-2 million tons of fresh water throughout the 168-day drilling operation. Cement plugs
were attempted to heal loss zones up to 3.2 km, but they were unsuccessful, and this solution was abandoned
(Friðleifsson et al., 2019). Interestingly, the successful drilling of this well and the management of the bottom hole
temperature may have been aided by the total losses. Above the total loss zone, the pipe would have been
insulated from the formation temperature by the empty annulus. The IDDP-2’s circulation loss issues highlight an
important challenge for SHR.
The St1 Deep Heat Project is an EGS-type direct heat project located beside an urban power plant in Otaniemi,
Finland (a district of Espoo), located 5 km from downtown Helsinki. The project aimed to reach a depth of 5-6 km
and supply 20-40 MWth of district heat to the city of Espoo from 100˚C fluid sourced from a stimulated fracture
network, re-injecting the fluid at 50˚C through an EGS doublet. While this is not analogous to an SHR project, this
case study provides rare insight into drilling to >6 km through hard rock, which will be required for most SHR
projects.
Initiated in 2014, St1 completed three wells in the Fenno-Scandinavian Shield by 2023. The pilot well (OTN-1)
penetrated 2 km, the seismic monitoring well (OTN-2) penetrated 6.2 km, and the stimulation well (OTN-3)
penetrated 6.4 km measured depth, with some tangential drilling for reservoir creation, making it the world’s
deepest industrial geothermal project to date. After creating the fracture network, OTN-3 underwent a 49-day
14
stimulation test that was temporarily successful despite considerable wellbore failure that required variations in
100-200 m long packer-sealed sections.
Figure 4 Geological diagram of the Otaniemi EGS site in Finland, including well location, depth, and 120˚C temperature horizon.
(I. T. Kukkonen et al., 2023a)
The geology of this area is complex, with cold granite rock, interbedded formations, and a geothermal gradient of
15-17˚C/km (I. Kukkonen & Pentti, 2021). The project’s drilling success proves the feasibility of drilling into hard
rock formations to unprecedented depths. Strada Global used air hammer drilling methods for sections of the
three wells (Figure 4), successfully drilling to 4,520 m for 72 on-bottom drilling days—the world record for drilling
depth into hard granite. However, drilling operations consumed 20,000 L of fuel per day. For the deepest well
(OTN-3), air hammer drilling was used from 300-3,300 m, followed by rotary drilling to reach 6,400 m (including a
deviation). The project experienced wellbore failure during the hydraulic stages, and a series of packers from
PackerPlusR were deployed at varying intervals to support the wells (Kukkonen et al., 2023b).
Like IDDP, Stn-1 also experienced casing failure at depth and fluid losses but with a significantly colder lithology
than IDDP. These common challenges indicate that the risk of circulation losses and borehole collapse likely apply
to most geological environments. While the Stn-1 project proves the potential of existing drilling techniques to
reach these depths, sustaining reservoir fractures at this depth remains an unsolved challenge. At ST1, the failure
to sustain the fractures at depth led to the eventual closure of their project (Kukkonen et al., 2023a).
Conclusions
The experiences of these frontier pilot sites indicate that existing drilling technologies are capable of:
15
These achievements are encouraging for SHR yet indicate that further engineering iteration is required to make
SHR geothermal projects technically and economically feasible. No single project has demonstrated enhanced ROP
into SHR through >4 km of hard rock. Further technology gaps may be discovered if the drilling challenges
experienced by recent projects are overcome; therefore, further in-field testing at sites with similar characteristics
is needed.
• Failure of downhole tools due to high temperatures, leading to inaccurate LWD or MWD,
• Failure of casing in corrosive or high-temperature environments, leading to borehole failure and collapse,
• Risk of blowouts due to high-pressure, high-temperature working fluids, and
• Circulation losses due to porous or fractured media.
Lastly, the Utah FORGE and Fervo projects have demonstrated impressive ROP (>25 m/hr) through hard rock using
specialized drilling technologies. However, Figures 2 and 3 indicate that the drilling performance is approaching a
mechanical limit, where improved drilling practices yield the same drilling rates as previous wells. As the wells
extend deeper to reach SHR in locations, this limit could affect the economic viability of SHR drilling. This limit may
be overcome with further innovation in drill bit technology.
16
3. Drilling system requirements for SHR
To handle the supercritical conditions of SHR (temperatures >374°C and pressures >22 MPa), a drilling system
requires additional capabilities in the following technology domains:
• Rock-destroying equipment,
• High-temperature downhole tools,
• Temperature management equipment, casing, cement, and corrosion inhibition, and
• Topside equipment.
Most SHR wells will be drilled through igneous and metamorphic rock with high compressive strength. The bit or
rock-destroying technology needs to deliver an ROP that is fast enough to make the well economically viable. If
the system uses conventional drill bits, roller cones, or PDC bits, the system needs to be designed so that the
weight on bit (WOB) and torque on bit (TOB) can be delivered to the cutter-rock interface at substantial depths
(>5 km) without suffering vibrational founder or high friction. If an alternative drilling methodology is selected,
such as percussive, waterjet, thermal spallation, millimetre wave (MMW), or plasma direct-energy drilling, then
the surface system/rig needs to be designed to deliver the energy necessary to fail the rock. Section 4 provides a
thorough analysis of state-of-the-art conventional, hybrid conventional, and direct-energy drilling methods,
assesses their TRL status, and provides innovative pathways to overcoming remaining technology gaps for SHR
drilling.
The bottom hole assembly (BHA) for an SHR drilling system must include downhole sensors to collect MWD and
LWD data. MWD tools provide direction and inclination data and enable the well to be steered to a desired
subsurface location. LWD tools provide data on resistivity, density, and porosity of the rock, while a drilling
mechanics sub collects data on pressure, temperature, 3-axis vibration, downhole torque, and downhole
WOB. However, most MWD and LWD tools have a maximum temperature rating of 175 oC, limiting drilling
operations to this temperature threshold. Downhole tools must be innovated to be deployed in high-temperature
domains for SHR projects to succeed. Section 5 provides a thorough analysis of state-of-the-art high-temperature
downhole sensors and monitoring systems, assesses their TRL status, and provides innovative pathways to
overcoming remaining technology gaps for SHR drilling.
An SHR drilling project requires a robust temperature and pressure management system, along with a well
control/integrity system that can manage hot fluids. The system must be designed to deliver a working fluid,
typically water, to the bottom of the well at total depth (TD) to cool the drill string and downhole electronic
17
components below the maximum temperature threshold of the tool or component with the lowest temperature
specification. An alternative drilling fluid that could be used for SHR drilling is supercritical carbon dioxide (sCO2)
controlled with an MPD system that maintains liquid CO2 in the drill pipe and gaseous CO2 in the annulus. Other
methods for managing the bottom hole temperature include insulated drill pipe, mud coolers, flow rates, and
continuous circulation. Section 6 provides a thorough analysis of state-of-the-art well temperature management
technologies, assesses their TRL status, and provides innovative pathways to overcoming remaining technology
gaps for SHR drilling.
While largely outside the scope of this study (see Bridging the Gaps: Well Design and Construction by
Suryanarayana et al., 2024), casing and cement are critical components of an SHR well. The drilling engineer
designing the casing and the cement program must consider several factors. First, the last casing string or last
uncemented hole section needs to be large enough to accommodate a sufficient flow rate to deliver the right
amount of heat/energy back to the surface. Casing string size has a significant impact on overall project economics.
Second, the drilling engineer must consider the material properties and the wall thickness of the casing so that
even after the casing strength has been de-rated due to the elevated bottom hole temperature, it still maintains
the mechanical properties required to deliver the heat back to the surface for the design life of the project.
The third key factor is corrosion inhibition. The casing needs to be made of a material that can withstand the
temperature and the chemistry of the fluids over the lifespan of the well. If the well is an EGS-style well, the
produced fluids are likely to be mildly corrosive and, therefore, the casing material needs to be able to handle
those corrosive fluids. If the SHR design is a closed-loop system, the chemistry of the produced fluid is likely to be
more benign and less prone to corroding the casing, so more standard materials could be used.
Lastly, the cement design is also critical to the long-term integrity of the well. The cement needs to have little to
no shrinkage or expansion at the borehole temperature. Various non-Portland cement designs have been tested
up to 350°C (Arbad et al., 2022) and, according to engineers from Halliburton (Interview, 2024), there are cement
formulations that will be able to deliver a strong casing bond at even higher temperatures. All the literature
supports a cement design that includes pumping cement back to the surface in all casing sizes to provide maximum
support to the casing, minimize casing expansion, and protect the casing from the thermal expansion of fluids in
the void spaces which can result in casing collapse. SHR cement will also need to be reverse circulated to ensure
the low-temperature retarding chemicals do not see SHR temperatures.
Further research is also needed to address the potential of formation creep or expansion at SHR temperatures
and pressures. If the formation (or some of the component minerals within it) behaves plastically at elevated
temperature and pressure, the last casing string may not need to be cemented in place. The formation would
expand, create a seal, and stabilize the casing in the hole. Cement and casing are discussed in detail in a separate
report in this series: Bridging the Gaps: Well Design and Construction (Suryanarayana et al., 2024).
18
3.5 Requirements for topside equipment
Reaching SHR requires wells with very deep true vertical depths, which must be considered before other drilling
factors such as rock hardness, high friction coefficients, and elevated temperatures. The rig must be sized to lower,
pull, and potentially rotate the heaviest string going into the hole. The substructure and derrick must be strong
enough to lower and pull the deep 9 5/8” (or equivalent) casing. For example, the weight of a 9 5/8”, 58.4 lb/ft,
P110 casing at 20,000 ft is over a million pounds. The rig derrick and top drive must turn the whole drill string in a
high-friction environment and provide adequate torque at the bit to fail the rock efficiently. Developers will need
a rig with some of the highest specifications available today (e.g., the Doyon Rig 26 in Alaska, which is drilling
exceptionally long O&G wells on the North Slope). We estimate the number of rigs with these specifications to be
less than 50 worldwide.
The control system and auto-driller (i.e., fully automated control system) are vital equipment for an SHR rig. These
electronic components optimize the control of drilling parameters and improve precision and control. When
drilling through extremely hard rocks, it is important to precisely control parameters on the drill bits. If a managed
pressure drilling (MPD) system is deployed, then it should be integrated with the rig’s control system (e.g., NOV’s
NOVOS™, Nabors, H&P, Precision and Patterson).
MPD systems are required to manage the pressures and temperatures at the surface and ensure safe operations
for the rig personnel. An MPD system is required for CO2 working fluid but may also be necessary in SHR with a
water-based system, especially if there is a risk of the water existing in multiple phases in the borehole.
Drilling SHR wells also requires a fit-for-purpose drilling rig and topside equipment that can power and integrate
all the considerations discussed in this section (ROP, high-temperature downhole tools, temperature management
equipment, drill string, casing and cement). For example, an MMW rig must have sufficient capacity to withstand
900 tons of force for the weight rating of the rig, and a 5,300-horsepower drawworks to raise or lower the metal
waveguide to depths >7 km. Rigs with these capabilities do exist today but are rare and difficult to source (Houde,
Araque, et al., 2021).
19
4. Rock-destroying equipment: Technology frontier and gap analysis
Rock-destroying technologies must overcome three core challenges to drill to SHR:
These challenges are inherently linked. To reach SHR, we must drill exceedingly deep wells into hard rock, which
would incur prohibitively high costs with existing drilling technologies. Improving bit performance and maximizing
bit longevity to reduce trip time are equally important for increasing ROP. In this section, we assess three
categories of drilling techniques, assigning each a technology readiness level (TRL):
1. Conventional rotary drilling (PDC, roller cones, and hybrid PDC/Roller cone systems),
2. Hybrid conventional drilling (percussive, waterjet, and particle impact), and
3. Direct energy (plasma and millimeter-wave).
The majority of conventional and EGS geothermal projects are drilled with a conventional rotary drilling methods
and some with percussive methods. Therefore, these mature drilling technologies are the most advanced
technologies for reaching deeper, hotter regimes. Major advancements in ROP have been demonstrated in recent
EGS-type projects in hard rock. Rotary methods, however, are originally designed to cut through soft lithologies
and must continue to be adapted and improved for geothermal projects in hot, hard rock. Percussive, waterjet,
and particle impact drilling are methods that specifically target the brittle state of basement lithologies and may
lead to further leaps in drilling efficiency.
As well depths increase, both rotary and hybrid conventional drilling methods will increasingly face challenges
from excessive string weight, high frictional resistance, and exceedingly long trip times, possibly negating
advancements in ROP generation. The advantages of direct energy methods—namely minimal downhole
equipment, few complications from high temperature and pressures, and less (or no) trip time—may ultimately
provide a better pathway to SHR. Direct energy methods are highly novel, however, and may not be
commercializable this decade.
Conventional rotary drilling techniques are the most mature and accessible drilling methods today. Most of the
equipment is readily available from vendors and drilling contractors. A major benefit of drilling with conventional
drill bits is the use of standard bottom-hole assembly design, drill string design, and top-side rig equipment.
Conventional rotary drilling techniques can be modified for SHR projects by optimizing the drill bit design and
enhancing the downhole assembly with heat-resistant materials.
20
This section explores four sub-categories of rotary drilling innovation:
• Efficiency (the optimization of bit and cutter design for hard rock drilling, control of downhole drilling
parameters and bit dysfunction mitigation to improve productive drilling time and reduce drilling costs),
• High temperatures (specialized drill bits for SHR temperatures),
• Directional drilling (methods and equipment for drilling deviated wells in SHR), and
• Performance in interbedded formations (bit designs and methods to overcome bit wear from
heterogeneous formations that can be encountered in basement volcanic sequences).
In traditional hydrocarbon drilling operations, fixed cutter drill bits with polycrystalline diamond compact (PDC)
cutters are typically used for lithologies of soft-medium hardness, while roller cone bits are used for hard rock
lithologies (Graham et al., 2017; Petty et al., 2020; Pink, Patterson, and Thoresen 2023; Self et al., 2021). Roller
cones, or tricone bits (Figure 5), comprise three rotating cones with tungsten carbide inserts mounted on each
cone. Applying weight-on-bit (WOB) causes the mounted inserts to impact and fracture the formation, which is
removed by the continued rotation of the cones that clear away the crushed rock (Roberts et al., 2022).
While roller cones are typically used for conventional geothermal projects, the rotating components of these bits
are prone to failure due to high frictional heating and high WOB and are limited by the temperature ratings of
elastomer seals and bearing assemblies (approx. 130˚C) (Petty et al., 2020). PDC inserts on tricone bits have been
used to improve bit life and ROP, but still proved less effective than fixed cutter bits (Self et al., 2021). Roller cone
bits rated to 300˚C were used at IDDP, but even though the design proved robust and reliable, the ROPs were still
only 2-10 m/hr—a consequence of the hard, volcanic formation.
Figure 5 Polycrystalline Diamond, Roller Cone and Hybrid PDC + Roller Cone bits. From the Baker Hughes Vulcanix TM Geothermal bit line
(Baker Hughes, 2024)
Fixed cutter bits have no moving parts and shear through rock formations. They are either driven from the surface
by the rotation of the drill string, driven downhole with a mud motor, or a combination of both. Like roller cone
21
bits, the cutters point-load the formation with applied WOB, crushing rock that is removed as the bit progresses
downhole (Self et al., 2021). Fixed cutter bit performance has been significantly improved in recent decades by
advancements in cutter grade and shape, with materials resistant to thermal abrasive wear enabling their use in
hard rock formations, including those expected at deep geothermal sites. These bits have been further engineered
to cope with supercritical geothermal settings, where elevated risks of thermal wear, mechanical damage, lateral
vibration, and excessive torque are expected (Roberts et al., 2022). Finite element analysis, laboratory
experiments, and in-field testing on axial impact and thermal wear have shown that V-shaped cutters outperform
round cutters in ROP by 16-57 percent depending on rock type, and round cutters failed before the V-shaped
cutters under thermal-abrasive and axial stresses (Rahmani et al., 2021).
An example of high-temperature, abrasion-resistant drill bits for geothermal drilling is the VulcanixTM line by Baker
Hughes. The VulcanixTM Tricone bit (Figure 5) features a patented all-metal sealing and bearing system, with a
stabilizing bit configuration rated to 260˚C. The tricone bit has been used in a series of wells in Turkey, some
reaching up to 300˚C, with a record performance of 3,680 m in 36 days. These wells are not SHR, not exceedingly
deep, and were drilled through quartz-schist type lithologies; therefore, performance may differ in other hard
rock formations. Baker Hughes’ VulcanixTM PDC bit was deployed at Utah FORGE, achieving a maximum ROP of 43
m/hr and an average ROP of 23 m/hr, drilling 34 percent longer runs than planned. Their bit is equipped with
ShockWaveTM shaped cutters that enhance durability and improve bit longevity and PrismTM shaped cutters that
apply a point load to absorb load and mitigate breakage under high WOB. The FORGE tests demonstrated high
ROP in hard rock, but the temperatures were <250˚C. Therefore, further design and testing for SHR conditions are
required. Lastly, the Vulcanix KymeraTM hybrid PDC and roller cone bit was used to drill through high-temperature
(<230˚C), hard, volcanic formations in Japan. Two wells were drilled (up to 50˚ inclination) in just 32 percent and
43 percent of the planned drilling time due to increased ROP performance. This bit should be deployed to hotter
and deeper wells for further performance testing.
NOV tested its novel PDC-cutter shape and bit at the Tauhara Geothermal Field in New Zealand. The bit was
specifically designed to withstand thermal and abrasive wear and front-face thermal overloading, using cutters
enhanced with stronger/denser diamond bonds, tougher carbide substrates, and high-pressure platforms
(Roberts et al., 2022). In-field performance of this bit (referred to as the “Steam-Trooper,”) was tested against
traditional PDC bits, tricone bits, and hybrid bit designs. It demonstrated a 16 m/hr ROP and superior bit longevity,
completing an interval of 650 m. The bit showed minimal wear after this run, so it was redeployed and completed
another 1,625 m run (1,027-1,541 m longer than the tricone bits and 773 m longer than hybrid bits). However,
these runs were achieved in sedimentary layers. The bit experienced highly erratic torque when it encountered
the hard-rock rhyolite formation (Roberts et al., 2022). While this bit has demonstrated impressive longevity, it
may require further innovation and testing for SHR and hard rock conditions.
The optimization of shape cutter and bit design for hard rock drilling is well documented and well tested. However,
the designs have not yet been run in SHR conditions. Further bit and cutter engineering may be needed to cope
with the plastic behaviour of rock at the brittle-ductile transition.
22
Pathways for further research and experimentation to overcome this uncertainty include:
• Finite element analysis of cutter and bit performance at super-critical depths and temperatures (e.g.,
Rahmani et al., 2021; Rasyid et al., 2021),
• Laboratory experiments in brittle-ductile conditions (e.g., Rahmani et al., 2021; Rasyid et al., 2021; Roberts
et al., 2022), and
• Field experiments in brittle-ductile conditions with state-of-the-art robust drill bits, such as the Steam-
Trooper or VulcanixTM Kymera.
Standard drilling optimization methods improve ROP and bit life using techniques such as real-time surface MSE
surveillance, testing bit/BHA design, drill bit forensics, and parameter mapping (Sugiura, 2021). At Utah FORGE,
mitigation against drilling dysfunctions, such as stick-slip and bit balling, was achieved by improving monitoring
systems and topside communications. It is well-documented that specialized training of drilling personnel on the
physics and signatures of dysfunctions significantly improves ROP and can result in up to 30 percent reduction in
total drilling time. Managers at Utah FORGE consider the two-day training facilitated by Utah FORGE to be one of
the key factors in the ROP improvements achieved at their wells. This training ensures that all drilling personnel
are uniformly informed of the causes and signatures of bit dysfunction and equips them with the diagnostic skills
needed to overcome these performance failures (Dupriest and Noynaert, 2022).
The balance of WOB and MSE management must be tailored to the specific context of the drilling project. In other
words, the strategy used to optimize these parameters at one site may not work at another. Other factors that
may impact efficiency and dysfunction management include operator fatigue, miscommunication, and varying
experience levels. These factors can lead to human error that may impact drilling performance. Further
experiments with automated or AI-assisted drilling systems with high TRL values could lead to improvements in
ROP. For example, the REVitⓇ ZT system from Nabors was successfully implemented at Utah FORGE to mitigate
stick-slip dysfunction (Figure 6). Similar software tools exist from multiple vendors, and all improve the
performance of the drilling system and reduce vibration.
23
Figure 6 WOB and RPM regimes conditions for stick-slip dysfunction for conventional versus REVitⓇZT drilling methods.
(Nabors, 2023)
Automated systems could include Corva’s AI-based ROP optimizer for predictive drilling, in combination
with Nabors’ SmartSuite. The SmartROSⓇ Rig predictive drilling system has been shown to increase ROP by 61
percent for neighbouring wells and is deployable on any rig. The SmartNAVTM automated directional drilling
guidance is designed to improve wellbore placement accuracy, has shown significant reductions in drilling time by
25 percent in comparative case studies, and is compatible with steerable or rotary steerable operations. The
SmartDRILLTM automated drilling system reduces vibrations and shock of downhill tools (reducing unplanned trips)
and the SmartSLIDETM directional steering control system optimizes cycle time.
To further improve WOB and MSE management, automated or assisted drilling rigs should be deployed at ongoing
geothermal sites, such as Utah FORGE, or future EGS experimental sites such as Newberry in Oregon and Chevron’s
project in California, both of which were awarded funding through the Department of Energy’s EGS Pilot
Demonstrations program in 2024. The MSE management and ROP of automated or assisted drilling rigs could be
directly compared to fully manual drilling rigs. Comparing MSE management and ROP between automated
systems and traditional manual drilling rigs will provide valuable insights into the potential benefits and areas for
improvement. This research will not only advance our technical knowledge but also guide future investments and
innovations in geothermal energy extraction.
For deep geothermal drilling, conventional drill bits must be adapted to cope with exceedingly high temperatures
(>300˚C). Thermal damage is typically exhibited at the cutter shoulders, where cutter shear length and sliding
distance are highest. Various round and V-shaped cutters have been engineered by NOV with denser diamond
grades that are resistant to abrasive wear. These cutters also achieve greater durability against thermal wear by
24
removing cobalt catalysts within the diamond at the cutting edges, increasing the diamond temperature rating
from 700˚C to 1200˚C (Roberts et al., 2021; Self et al., 2021).
The dynamics of the fluid channeled through the top of the bit also play a critical role in regulating bit temperature,
minimizing thermal abrasion, and cleaning the cutters. The main parameters affecting hydraulics include cooling
fluid flow rate, nozzle configuration (Figure 7), mud weight, total flow area, and mud type. These parameters can
be optimized to distribute the correct amount of cooling to the most vulnerable cutters (Roberts et al., 2021).
Figure 7 Nozzle configuration to optimize drill bit cooling with fluid flow
Left shows the placement of 10 nozzles on an optimized bit, center/right shows the flow distribution and velocities from six nozzles on
two bits (Roberts et al., 2021).
As discussed above, the Baker Hughes VulcanixTM tricone and hybrid bits are rated to 177˚C, 288˚C, and 300˚C
using all-metal cone seals, high-temperature rated lubricants, and upgraded bit metallurgy. These bits underwent
laboratory experiments with 300˚C water-based drilling mud, 10,000-30,000 lbs (4,500-13,600 kg) WOB, 90-180
RPM, and 6000 psi into Sierra White granite. There was no loss of grease, and seals/bearings retained their
diaphragm depth, indicating no loss of lubricants that would lead to frictional wear. Drilling engineers have also
observed that carbide outer compacts in the bits had minimal wear (Petty et al., 2020; Stefánsson et al., 2018).
These bits were later tested at the Iceland Deep Drilling Project (Section 2.2) and demonstrated an average ROP
of 4.2m/hr and 3.4 m/hr for the 300˚C and 288˚C bits respectively. Hybrid and conventional bits were also used to
deepen the vertical section of the well. Meanwhile, coring bits, along with rotary, motor-driven and metal-to-
metal (M2M) BHA, were used at various stages. The highest ROP achieved was 10.3 m in run five by a 300˚C roller
cone under 121˚C conditions, but this was only sustained for 69 m due to circulation losses and the deployment
of a coring program that was unsuccessful (see Stefánsson et al., 2018 for a full review of bit performance per
run).
Several drill bit technologies can withstand 300˚C; however, the remaining 75-100˚C expected at supercritical
temperatures must still be “unlocked.” It should be noted that drilling through 400˚C rock does not mean that the
bits will encounter 400˚C due to significant cooling from the drilling fluid. The PDC cutters may encounter 400˚C
but are rated 800˚C. Access to these extreme temperature conditions remains a major challenge for
experimentation. Some potential solutions include:
25
• A global consortium of HTHP labs to conduct necessary R&D experimentation and coordination between
industry partners with innovative technologies to co-develop the drill bit technologies with supercritical
temperature capacities;
• On-site experimentation at high-temperature locations, specifically involving sites studied at the Japan
Supercritical Project, Newberry Volcano, and the Iceland Deep Drilling Project; and
• Thermal analysis for bit design with software such as ReedHycalog up to 450˚C (as their highest
temperature tests are only up to 250˚C).
Precise directional drilling tools are required to deploy the lateral well sections that can operate in >375˚C and at
great depths. The borehole must be placed in a precise location to either enable a well-to-well intersection or to
optimize lateral borehole separation for reservoir creation and stimulation. Both cases require deviations from a
sub-vertical well. Therefore, directional drilling technologies must also reach SHR capacity.
Directional drilling requires a drive system that can be steered in three dimensions. Using traditional technology,
three-directional steering is performed with a mud motor and/or a rotary steerable tool. Both these drive systems
have temperature limitations of between 175˚C and 200˚C, thus the drilling system will need to be cooled below
these temperatures. Companies like Hephae are pushing the temperature limitations of both rotary steering
systems (RSS) and MWD to above 200˚C (Section 5). The improved efficiencies from an RSS system were observed
at the Utah FORGE’s production well 16B (78)-32, that was inclined at 65° parallel to the injection well (16A (78)-
32). The production well achieved a much higher ROP due to increased bit aggressiveness and the implementation
of a rotary steering system (RSS) in the inclined sections of the well. The primary improvement from the RSS was
the hole smoothness (Figure 8), which drastically reduced frictional drag on the drill string and the integrity of
other down-hole components such as packers (Noynaert, GRC 2023).
Other important considerations for directional drilling emerge from the nature of the rock itself. To access SHR,
wells must be steered through hard basement formations that have high friction coefficients. Taking the rock type
into consideration, the assembly must be designed to minimize vibration. Tools like roller reamers must be run in
the BHA, and the drill string must be as stiff as possible to minimize torsional vibration and maximize weight
transfer. To reach the vertical depths needed for SHR (5-10 km), it is essential that drilling is executed with the
lowest number of doglegs and the lowest tortuosity possible (otherwise the rig may run out of available surface
torque), which may require more investment in RSS tools to maintain verticality.
In the last hole sections of SHR wells, lubrication will become an important part of directional drilling. The ability
to steer the wellbore and minimize vibration may require the addition of lubricants to the drilling fluid. Research
into higher temperature variants of these lubricants is needed. Graphite can be used, but it is expensive and can
reduce reservoir permeability (so it should not be used in the final hole section of an EGS well).
Directional drilling an SHR well above 374˚C is possible today but, depending on the efficiency of the temperature
management system, it may be necessary to complete all the directional work with electronic tools before the
26
highest temperatures are encountered. The modelling work done on insulated drill pipe suggests that directional
work can be done at depths of 8,000 m and with formation temperatures >374˚C, but a set of full-scale test joints
needs to be produced and tested, followed by the development of a complete string to validate that the bottom
hole circulating temperature can be maintained below the operating temperature maximum of the tools.
Deep SHR wells must also consider the surveying program. To be able to accurately hit a small target or maintain
exact proximity to another well, the survey program may require running gyro tools at the end of each section
and advanced survey techniques while drilling. In SHR CLG wells, magnetic ranging tools will (discussed in Section
5.1) be required to make a well-on-well intersection; these tools are currently being tested in Geretsried, Germany
by Eavor (but not under SHR conditions).
Figure 8 Profile of curved section of well 16B, showing correlations of hole smoothness and rotary-motor systems used
(Noynaert, GRC 2023)
SHR conditions are generally found in basement rock at depths >5 km. Engineers at Fervo Energy have stated that
lithological hardness becomes more uniform with increasing depth, and current drill bit technology will not be at
27
risk of premature bit-wear from drilling through heterogeneous formations, such as sediments, interbedded tufts,
calcites, and granite. In heterogeneous media, one ReedHycalog bit managed to drill through sediment,
interbedded tuffs, calcites, and granite, demonstrating that conventional bits are sufficient in interbedded
domains. Conversely, experts at Utah FORGE consider interbedded formations to be a first-order innovation gap
for hard rock drilling, and the Stn-1 project in Finland encountered complex geology in the Fenno-Scandinavian
shield, a geological regime with presumably uniform rock hardness.
When PDC cutters transition from soft to hard formations in heterogenous domains, the drill string rotation is
slowed, yet the force on the bit remains the same (Graham et al., 2017). The excess force dissipates through the
cutters, causing breakage or chipping if the force exceeds their fracture strength. Tricone bits may be better suited
for these conditions, as they penetrate the rock with the weight of the drill string rather than the rotational force
applied by fixed cutters (Graham et al., 2017; Rahmani et al., 2021).
Alternatively, Schlumberger has innovated a conical diamond element (CDE), fixed with PDC cutter teeth (Figure
9) that improve the implementation of fixed cutter bits in hard, interbedded rock. CDE teeth are 25 percent more
resistant to wear, reduce vibrational founder on the bit, and have 50 percent more impact strength and 50 percent
more diamond thickness than PDC teeth. Where PDC cutters shear through rock, CDE cutters “plow” through the
formation, and their position behind the point of maximum tangential velocity experienced by the PDC cutters
protects them from impact damage (Graham et al., 2017). They also use 26 percent less torque than PDC bits,
improving directional capability due to the reduction of torque fluctuations at the tool-face (Rasyid et al. 2021).
Rasyid et al. (2021) performed finite-element analysis of CDE bits of varying sizes before Schlumberger drilled a
test well at the Blawan Ijen geothermal field in Indonesia. Their simulations tested drilling andesite, tuff, and
brecciated volcanic formations (with 6, 7, and 8-bladed CDE bits in comparison with tricone bits) for 17 ½”, 12 ¼”,
and 9 ¾” well sections. The CDE bit consistently improved stability (less vibrational founder) and ROP compared
to tricone bits. In-field drilling in similar rock types and well pressure conditions for the Blawan Ijen test wells
resulted in an ROP range of 5-30 m/hr, saving 6-9 days of drilling compared to a tricone bit (Rasyid et al., 2021).
Furthermore, the CDE bit was tested in heterogeneous lithologies at The Geysers and achieved 4-8 m/hr ROP—a
28
45 percent increase from the PDC bit, with the potential to be further improved (Graham et al., 2017; Song et al.,
2023).
The TRL of fixed cutter drill bits equipped to excavate interbedded formations is 9, as demonstrated by the ROP
tests at the Blawan Ijen geothermal field. However, these bits have not yet been tested under supercritical
conditions, and are not yet rated for 400˚C. Recommendations for further experimentation and testing include:
• Continue to use Utah FORGE as a site for research on overcoming interfacial severity (specific methods
include adaptive drilling methodologies that can detect fractured lithology before intersection, potentially
through acoustic and sonic techniques that can be incorporated into automated drilling methods),
• Deploy CDE bits in deep, cold, heterogenous geologies (e.g., Stn-1) to compare performance to
conventional rotary bits,
• Conduct in-field tests or experimental bench tests in a laboratory with SHR capacity, and
• Create the equivalent of Utah FORGE in a volcanic geological environment featuring igneous rock with a
high variability of hardness.
Conventional drilling is the primary method for accessing geothermal resources in hard crystalline formations,
with some cases in supercritical temperatures. Individual bits can be specialized for extreme temperatures,
interbedded formations, or extended bit life; however, no single bit features all these specifications. Combining
these features may be the most efficient way to accelerate conventional drilling for SHR applications. However,
this innovation requires coordination between major drilling firms.
The main obstacles to coordinating efforts between major drilling firms are the reluctance to share trade secrets
and intellectual property. However, there are ways around this. For example, NOV and Schlumberger have
demonstrated the ability to collaborate through licensing agreements. To propel this technological frontier
forward, we need to develop creative approaches for incentivizing collaboration while also protecting intellectual
property and ensuring the profitability of firms.
The other major gap for drilling into SHR with these state-of-the-art technologies is the lack of accessibility to
supercritical conditions. There are no known laboratories capable of testing these drilling technologies at >374°C
and testing in-field with these conditions is expensive and high-risk. Following extensive experimentation in a
controlled laboratory setting, high-temperature, SHR-rated drill bits and BHA should be deployed in active SHR or
EGS projects for further optimization of deep, hot rock drilling. The Stn-1 wells in Espoo, Finland, which have
ceased commercial production of direct heat, may offer a unique location to verify ROP performance at extreme
(>6.3 km) depths in hard rock.
The Newberry Volcano or IDDP project can offer access to extreme downhole temperatures and are ideal locations
for experimenting with high-temperature bits and tools, insulation technologies, and alternative drilling fluids such
29
as sCO2 (see Section 6.2). The Japan Beyond-Brittle Project also strives for supercritical depths and temperatures
and may offer the chance to experiment with these various drilling technologies.
Further experimentation on drilling into supercritical conditions could be facilitated by the Utah FORGE
demonstration site (with consent from the DOE), and by recipient projects of the recently announced DOE
Geothermal Earthshot funding (Mazama Energy at Newberry Volcano, Chevron in Sonoma County, and Fervo
Energy in Milford, Utah). For drill bit design, modelling, and rig servicing products, NOV, Nabors, and Schlumberger
have facilities and a breadth of expertise to engineer tools to overcome specific technology gaps, in partnership
with specialized research entities and universities.
While generally not as mature as conventional rotary drilling methods, hybrid conventional methods may
ultimately provide the optimal pathway to accessing SHR. Percussive and waterjet drilling were initially designed
for unique applications in hydrocarbon exploration and have been adapted for deep, hard rock drilling.
Meanwhile, particle impact drilling was developed specifically for hard rock drilling. Most hybrid conventional
BHA’s are combined with rotary drill bits, with the percussive/waterjet/particle impact modifications designed to
target the brittle state of hard rock formations to improve drilling performance and increase ROP. For example,
waterjet methods alter the stress state at the bottom of the hole to improve its failure, and percussive and
particle-impact fracture the brittle hard rock.
The same core challenges facing conventional drilling apply to hybrid conventional drilling methods, namely:
efficient drilling through deep, hard rock formations under exceedingly high temperatures and pressures. The TRL
status of percussive drilling is 7, while waterjet drilling and particle drilling are at a TRL status of 4-5 for superhot
hard rock applications that may be advanced with further in-field testing. The advantages of hybrid conventional
drilling may prove superior to purely rotary techniques in the pursuit of SHR, although further experimentation in
SHR conditions is required for all methods.
Percussive (i.e., hammer) drilling is equal in maturity to rotary drilling methods and has demonstrated promising
performance in hot, hard rock drilling for SHR. A typical percussive BHA consists of a drill bit, a piston, and hammer
casing. The piston pneumatically impacts the bit repeatedly into the rock, with the casing connecting the hammer
to the drill string (Depouhon et al., 2015; Song et al., 2022). Percussion drilling has produced improved hole
geometry, reduced stress on the drill string, improved bit longevity due to reduced contact time with the rock,
and produced better cutting shape and size for transport. Percussive drilling can, however, lead to wellbore
instability due to excess energy transfer in soft formations, poor performance in soft rocks, and vibrational
dysfunction, and can have uncertain optimal hammer frequencies (Bruno, 2005).
Percussive drilling is commonly used for hydrocarbon exploration and has demonstrated significant ROP: >24 m/hr
in soft rock formations, and up to 20 m/hr in hard rock formations at 6,000 m depths. Fervo has demonstrated
30
that percussive drilling methods can maintain a straighter well path relative to PDC bits (because of the lower
WOB). Several firms are currently adapting percussive drilling methods to reach supercritical temperatures. UK-
based Strada Global and HydroVolve systems both combine rotary and percussive methods, and the EU Horizon
2020-funded Orchyd project combines Drillstar Industries’ MUDHammer with a waterjet system.
Alternative pneumatic systems for percussive BHA have been investigated by Lehmann and Reich (2015) to
optimize very deep, hard rock drilling. These systems include mechanical, thermo-mechanical, hydraulic, and
electromagnetic drives. Numerical experiments on percussive drilling up to 300˚C have shown that the tensile
strength of rock decreases with high temperatures (Song et al., 2022). Therefore, SHR may actually enhance ROP
for percussive drilling.
Strada Global
Strada has a fluid hammer operating system that combines rotary and percussive drilling designs. It applies a
dual-fluid system, using fresh or saline water to drive the hammer, which reduces corrosive wear on the bit. In
Australia, Strada’s fluid hammer was used to reach depths of over 6,000 m, achieve 20 m/hr ROP in 200 MPa
hard rock including granite and basalt, reduce drilling costs by 70 percent, improve bit longevity, and drill in
underbalanced and overbalanced conditions (Olijnyk, 2024).
HydroVolve
HydroVolve’s GeoVolve HYPERDRIVE system (Figure 10) is a hybrid rotary and percussive drilling method, designed
specifically to drill deep, hot, hard rock wells for geothermal projects. A rotating PDC bit drills with the same
mechanics as a conventional drill system, but the BHA is augmented to deliver a cyclic axial impulse that pre-
fractures the rock, reducing its compressive strength and decreasing the rotary shear required to clear the rock.
This system is purely mechanical; therefore, it is not prone to electronic failures due to high temperatures. Specific
capabilities of the HYPERDRIVE system include the ability to: (1) operate up to 300˚C; (2) be powered by the
pressurized flow of any drilling fluid, from freshwater to high-density drilling mud; (3) maintain typical hydraulic
systems; (4) maintain circulation density and sufficient wellbore pressure; and (5) protect the BHA from impact
force damage.
31
Figure 10 Schematic of the HydroVolve GeoVolve HYPERDRIVE system
(Moyes et al., 2023)
The system is compatible with a conventional rotary BHA. The drilling mud hydraulically drives a high-frequency,
percussive axial impulse system transmitted through the bit, fracturing the rock cleared by the rotary drilling
mechanism. Designed in partnership with ZerdaLab, the bit is optimized to cope with axial impulse load, using a
large shoulder radius and CDE-type cutter configuration to maintain lateral stability. It has no elastomeric seals
and is thus ideal for SHR conditions.
Full in-field and laboratory testing has been conducted using the HydroVolve assembly at their onshore Drilling
Test Facility, using a rig with downhole monitoring of WOB, axial impulse load, pressure, ROP, RPM, and torque,
with data transmitted at 10,000 Hz. This initial testing was followed by a series of trials in Hungary achieving an
ROP of 2 m/hr for a total depth of 229 m, sustaining downhole temperatures of 220˚C in interbedded sedimentary
formations. This equipment was also used to drill 5,800 m through interbedded sediments in Ukraine sustaining
2 m/hr (HydroVolve Case Study Documentation, 2023). Further testing in hard, SHR conditions is required and
ROPs of greater than 2 m/hr will need to be demonstrated to be competitive with other technologies. Field trials
in New Zealand are scheduled to take place in 2024.
Orchyd Project
The EU Horizon 2020-funded Orchyd Project, in collaboration with Drillstar Industries, developed a hybrid high-
pressure waterjet and mud hammer. This hydro-mechanical drilling system operates to reduce the confining
stresses of the rock at the drilling front, improving drilling efficiency and reducing the impact of hard, abrasive
formations on bit-life. The release of stress is achieved by grooving the rock at the drilling front with the high-
pressure water jets, causing the unloading of compressive stresses and promoting fracture propagation to reduce
rock strength (Wang et al., 2021). The Drillstar MUDHammer, equipped with a diamond-insert reinforced kerf-
shaped drill bit (Figure 11), then impacts the weakened rock at high frequencies, generating cuttings that are
cleared away with a circulating drilling fluid. The axial vibrations of the drill string also pressurize the water,
reducing vibrational dysfunction and conserving energy.
32
Figure 11 Orchyd and Drillstar hydro-mechanical drill system, and the bit-jet-rock interaction at the drilling interface.
(Orchyd, 2023)
With these designs, current system testing at the ARMINES laboratory facility in Pau, France demonstrated an 80
percent increase in ROP with the combined system, a rate of 4-10 m/hr (Jahangir et al., 2022). With more planned
updates on the way for design and validation, there is potential for this number to increase. The target of the fully
completed system would be an ROP of roughly 20-25 m/h in hard rock. This drilling approach, however, has not
been extensively tested in ultra-high temperature and pressure conditions.
Percussive drilling, both with air and mud hammers, is at a TRL status of 7 for deep drilling into hard SHR. Several
established firms are developing specialized BHA for deep geothermal drilling, and several frontier geothermal
projects have invoked percussive drilling systems into their drilling program, notably Stn-1, Fervo, and IDDP-2.
Drilling into SHR is within reach. The main challenge is access to locations or facilities with HTHP conditions.
Percussive drilling methods are likely to encounter challenges in SHR conditions. Air hammer drilling will not
perform efficiently at great depths due to well instability and loss of force to the drill bit, and mud hammer drilling
may incur substantial water loss if operating in porous or fractured media. A collaborative research project
between Mines ParisTech and Drillstar Industries has identified four focus areas for how to adapt percussive
drilling for deep geothermal projects (Gerbaud et al., 2022):
33
• Theoretical development, computational analysis, and experimentation on the stress regime at the drilling
front in a deep borehole for optimizing cutting elements and bit configuration,
• Cutting element optimization at the rock-bit interface under specific drilling conditions,
• Design of a hydraulic hammer for 8 ½” boreholes, and
• In-field testing.
The MINES Paris Tech and Drillstar Industries study successfully designed an 8 ½” drill bit with a 300 percent ROP
increase relative to conventional percussive systems in a hard rock bench test, demonstrating significant
improvements in vibrational dysfunction. They have not, however, experimented in SHR conditions nor conducted
substantial in-field testing. One approach to advancing their technology is to partner with research groups with
the capacity to test at supercritical conditions, or field sites with extreme HTHP in-situ conditions, such as IDDP
(like Strada Global’s drilling program at the Stn-1 Deep Heat project). The Department of CNPC Engineering
Technology in Beijing has also conducted extensive modelling and laboratory experiments on stress wave
propagation mechanisms and rock fragmentation up to 300˚C (Song et al., 2022). Collaborations between these
groups could advance the modelling of HTHP conditions.
Percussive systems have achieved 25 m/hr ROP and reached supercritical pressures and temperatures. While
rotary drilling systems have demonstrated superior ROP performance, percussive methods may achieve similar or
better rates with further optimization for SHR and deployment in field settings. Suggested next steps include
international collaboration to deploy SHR-rated percussive systems at in-field demonstration projects to further
optimize this technology.
Waterjet drilling uses high-pressure water to remove rock formations to create a borehole. This method weakens
the formation within and surrounding the borehole, reducing the strain on the drill bit. There are two applications
for waterjet-based drilling: (1) drilling lateral passages and (2) vertical drilling. Lateral passage creation (e.g., radial
jet drilling) has only been used in soft lithologies for stimulating hydrocarbon reserves or conventional geothermal
systems to increase their output. Vertical wells are drilled with either continuous high-pressure streams, pulsed
streams, or through a combined system utilizing a conventional bit with jetting attachments.
Current system-level performance testing completed by the Orchyd and ThermoDrill projects shows a >30 percent
reduction in stress within the rock formation at a rate of 4-10 m/hr ROP at 65 percent reduced overall cost (Orchyd,
2022) and a ROP increase of 50 percent at 30 percent of the cost (Stoxreiter et al., 2018) relative to conventional
drilling methods. While these systems and designs have demonstrated significant performance increases, the full-
scale technology requires further testing. Top-side equipment must be designed to support the extreme volumes
of fluid delivered down hole (600 L/m in the case of the Orchyd Project), especially at extreme depths. Maintaining
that pressure to depths of up to 10 km will be critical for the success of waterjet designs.
34
Figure 12 Breakdown image of a radial jet drilling system within a borehole.
(Left) (Kaldal et al., 2019), Drillstar’s MUDHammer piston-based pulsed water drilling head (Center) (“Drillstar - MUDHammer,” n.d.), and
conventional drill bit with the addition of two high-pressure waterjet heads (Right) (Stoxreiter et al., 2019)
While these developments have been tested in a conventional geothermal setting, further validation and
development is required for SHR drilling. In the case of radial jet drilling, some key areas of design improvement
were identified by Ahmed and Teodoriu (2023) at the Stanford Geothermal Workshop in 2023:
Overall, waterjet drilling is most effective when combined with another drilling system. It increases overall ROP
and enables a lower frequency of tripping to surface. The reduction of downhole stresses is one of waterjet
drilling’s biggest advantages. However, the technology requires more validation based on the standoff distance
to the rock surface to maximize stress reduction. Surface systems need to be evaluated to ensure consistency
with significantly increased flowrates at long distances through the drill string.
Particle drilling involves the injection of steel shots of various sizes into the drilling fluid through the drill bit. The
impact of accelerated drilling fluid bearing the steel shots weakens the formation to improve efficiency of the PDC
drill bit. Increasing the run length per bit will improve the economics of SHR drilling by improving bit longevity and
reducing trip time.
Particle drilling is a process whereby steel shots are injected at pressure into the main flowline that runs from a
rig's pumps to the drill string. The steel particles then flow down the drill pipe into the drill bit and accelerate to
500 ft/sec. The drill bit rotates at roughly 60-80 RPM, and the particles cut a bottom-hole pattern in the rock (Pink
et al., 2022). The remaining half inch of de-stressed rock is then removed by the PDC cutters in the reaming section
of the bit (Figure 13).
35
In January 2021, the company Particle Drilling (PD) and NOV partnered to design a hybrid particle impact and PDC
bit and conduct lab and field testing in hard igneous rock in central Texas. This trial demonstrated significant
longevity improvements relative to traditional PDC bits, drilling 990 ft in extremely hard (45,000 to 50,000 psi)
pink granite at 45 ft/hr for 990 ft (whereas a single PDC bit was worn after 122 ft). The PD/NOV bit was tested at
Utah FORGE in 2023, with the aim to drill a 9 ½” hole section 500-600 ft in one run at 75 ft/hr, and capture and
recycle all the shots. The two drilling runs were unsuccessful, with penetration depths of less than 70 ft, a result
of the nozzle configuration on the hybrid drill bit. Based on the post-run analysis, the PD/NOV team developed a
new hybrid bit, specifically designed for Utah FORGE’s granite. This technology has potential for generating high
ROPs and completing long bit runs as the steel shots can be continuously replaced at the surface. With further
runs and iterative design improvements, this may result in less trips and reduced costs.
A significant technology gap is compatibility with MWD. To drill directional wells efficiently, particle drilling
requires an MWD telemetry system that can send signals uphole without a pulser or modulator. Currently, when
putting a conventional MWD tool in the BHA, the steel shot will damage the modulator or pulser. Conversations
with Halliburton suggest that there is potential for the Halliburton negative pulse tool to be modified to handle
steel shots. Wired pipe could also be used as a telemetry method if the circulating fluid temperature is kept below
150°C.
Canopus Drilling Solutions (CDS) in the Netherlands is also developing a hybrid particle impact and PDC system
(Interview with Jan Jette Blangé, Feb. 2024). Additionally, they are designing and testing a complete MWD package
and steering unit; the complete system is called Directional Steel Shot Drilling (DSSD). The Canopus system’s basic
principles are the same as PD: it uses high-velocity steel shot on the hard brittle formation. The technologies differ
in that Canopus uses the shots to steer the well and uses a PDC bit to remove the weakened rock, whereas PD
solely reams a small section of destressed rock on the outside of the hole.
The Canopus system underwent significant testing and validation between November 2022 and June 2023 at the
Rijswijk Centre for Sustainable Geo-Energy (RCSG) of TNO in the Netherlands (Knebel et al., 2023). The goals of
these tests were to assess the performance and feasibility of individual subsystems and provide evidence of the
complete integration of DSSD technology into an efficient operational control system. The testers were able to
36
mitigate risks with the sub-surface control system and improve their control of the steel-shot injection unit. A
detailed description of this testing is described in Knebel et al. (2023).
Figure 14 A picture of drilling rig VERMEER 55 in the tunnel at VersuchsStollen Hagerbach (VSH)
(Knebel et al., 2023)
The system was then taken to the VersuchsStollen Hagerbach (VSH) gallery tunnel in Switzerland where a
horizontal rig was used to drill 2,125 m sections (Figure 14). The testing at VSH demonstrated that the addition of
steel shots tripled the drilling speed in soft rock. Despite some challenges, the performance of the downhole
control unit met expectations and operators demonstrated that they could adequately clean the hole.
PD’s hardware is not affected by the temperature and pressure of drilling through SHR; however, the injection
process needs to be closely monitored when making drill pipe connections. PD’s particle impact drilling is market-
ready but requires further engineering, bench testing, modelling, and in-field operations to improve performance.
Testing at the Utah FORGE site or at a similar open access, hard rock drilling experimental site could accelerate
the development of the technology. PD’s particle impact drilling is at TRL 5, as it has been run multiple times at
full scale in the field but has remaining technology gaps.
Canopus is pushing towards market readiness in 2024 but must perform a full-scale drilling test with a vertical-
mounted conventional rig. Its tests thus far have been in soft rock and so it must validate its ROP improvements
in hard igneous rock. The system is designed for 6” hole sections (smaller than a typical geothermal well), although
the system can be scaled up without much challenge. Full-scale testing with good offset PDC data will provide a
good benchmark on performance. Another concern is the Canopus MWD system is an electromagnetic (EM) tool.
EM tools have the advantage of no moving parts in the fluid flow; however, they have major signal attenuation
37
problems in deep hard rock and formations with high conductivity. The report from the testing at VSH mentioned
signal challenges. EM tools were not used at Utah FORGE due to insufficient signal transfer. Halliburton’s negative
pulse MWD may be compatible with the Canopus system.
Canopus needs full-scale testing on a commercial project to determine whether the steel shots erode the PDC drill
bit. The images of the bit from VSH showed significant erosion, but it is not clear whether that erosion is from the
formation or the steel shot (erosion and impact damage were also seen on full-scale tests conducted by PD). The
authors of this report believe the modelling and testing that Canopus published on hole cleaning is thorough and
the steel shot can be cleaned and circulated out of the hole in horizontal sections. Canopus should form a
partnership with one of the leading drill bit manufacturers to conduct advanced fluid dynamics modelling with the
inclusion of particles. Canopus’s DSSD system is at a TRL level 4 and needs full-scale testing on a commercial
project or test rig in hard rock.
Direct energy drilling involves the destruction of rock by exposing it to high-powered energy sources, such as
plasma or millimetre-wave (MMW) energy. This focused energy weakens or vaporizes the rock. Cuttings are then
cleared by a circulating purge gas or drilling fluid, exposing the next segment of rock to further direct energy,
deepening the well. Direct energy drilling systems have several advantages over conventional drilling, including:
Direct energy methods perform best in hard rock conditions. Conventional drilling methods would still be required
to drill through the top 2-3 km of sedimentary overburden to reach basement rock before deploying the direct
energy system. The process of switching from one drilling system to the next is under development, and further
innovation may be required to ensure a conventional rig is compatible with a direct energy rig. Direct energy
drilling for rock excavation is relatively novel, therefore, these methods have low TRL statuses: TRL 3-4 for plasma
and TRL 2-3 for MMW.
The driving mechanism behind plasma drilling technologies is the application of electrical currents through rock
formations to create a plasma pulse through the grain boundaries within the rock, which can be achieved by direct
contact to the rock interface or while maintaining a stand-off distance (Figure 15). Both methods of plasma drilling
rely on an electrical charge to deteriorate the boundary layers as it travels through the rock pores. The
performance is dependent on the electrical field of the dielectric drilling fluid and the geological formation on a
micro timescale. When the discharge(s) of the pulses are less than 500 ns, the electric field of the fluid is greater
38
than that of the solid rock. The rock deteriorates before the water, resulting in the breakdown of the formation
(Q. Zhang et al. 2023). Figure 16 shows the application of plasma drilling to a rock sample using a Zaptec drill bit
(Agrawal et al., 2016).
Figure 16 Schematic and real demonstration of plasma drill bit drilling through hard rock
(Agrawal et al., 2016).
With a projected consumed energy of 100-200 J/m3, plasma technology requires roughly 3-5x less energy than
conventional drilling systems. The low energy requirements reduce costs by an estimated 17 percent, with the
possibility of up to 90 percent reductions with future improvements. Since this method can be completed without
the use of conventional drilling bits, the need for WOB is removed.
Slovakia’s GA Drilling is on the forefront of the application of plasma systems for geothermal drilling. They
designed a system that can be integrated with existing conventional drilling methods (Anchorbit®), as well as a
standalone design for ultra-deep depths (Plasmabit®). GA’s Plasmabit® system hopes to offer several advantages
over conventional systems—namely, increased ROP in hard rock formations, reduction of downtime due to
mechanical component wear, and reduction of noise generation and resonance in deeper formations. Testing of
39
Anchorbit® was completed by ETH Zurich and demonstrated a high ROP of 7.2 m/hr in granite samples. The
combination of conventional methods and plasma drilling alleviates the challenges of conventional systems,
including tooling wear, tripping time, and low ROP in hard, granite formations. Nabors, who are collaborating with
GA on plasma drilling R&D, highlighted the additional challenge of needing to switch out fluid types when
switching from a conventional drilling system to drill the sedimentary overburden, to a plasma drilling system for
the hard basement rock (Interview with Nabors, February 2024).
According to Liam Lines (Interview, March 2024), GA Technologies is developing another hybrid conventional-
plasma system using the shape and geometry of a conventional drill bit with the strategic placement of nodes for
the electrical pulses to follow. Through their proposed geometry, the pulsed plasma will weaken the rock
surrounding the borehole. By modifying the locations of the cutters, the pulses can interact with the formation,
reducing stress and damage on the tooling. This hybrid system is set to be tested in approximately 18 months at
Nabors’ facility in Texas.
While plasma drilling technologies show promise for reaching SHR, technology gaps remain. For plasma drilling of
hard rock formations (especially granite), the breakdown of the rock formation plays just as critical a role as the
technology used to remove the material. The variance within different granite formations plays a significant role
in the amount of energy consumed by the drill system and the amount of material removed during each pulse. A
recent study (Walsh and Vogler, 2020) shows a significant relationship between the pore size and the amount of
total energy consumed during the operation, with implications for ROP and the total required energy to complete
the borehole.
While controlled experimentation improves the understanding of these behaviors, larger-scale experimentation
is necessary. If the performance of plasma technologies is verified, the overall reduction in required power could
be in the range of 100-200 J/m3 as opposed to the 600-950 J/m3 required by conventional drilling technologies
(Walsh and Vogler, 2020).
Another key factor in the success of plasma drilling is the drilling fluid. Most geological-based laboratory tests use
water as the dielectric fluid within the system, but engineers need to better understand how to maintain clean
drilling fluid downhole in field conditions. GA’s hybrid conventional plasma technology shows the most promise
but must overcome the challenge of maintaining clean dielectric fluid and standard drilling fluids. Engineers from
Nabors acknowledge this issue requires further work, focusing specifically on where cuttings are removed and
difficulties when switching between fluids.
Another focus area for plasma drilling is the delivery of power to the drill bit, which should be considered in the
context of the existing directional control systems of drilling platforms. A more seamless integration with current
drilling systems could reduce the overall complexity of plasma drilling.
40
Despite these remaining technical challenges, plasma-based technologies show promise when compared to
relying solely on conventional drilling. With reduced overall energy requirements, competitive ROP, and reduced
cost due to tooling and time on task, plasma techniques may become a key technology for SHR drilling.
Millimeter wave (MMW) drilling is a direct energy drilling method that vaporizes rock by applying intense
electromagnetic energy to the rock surface, powered by a gyrotron and transmitted downhole by a metal
waveguide. Vaporized debris is excavated from the well by a circulating purge gas. MMW drilling is designed for
hard rock conditions such as the crystalline basement, therefore conventional drilling must first be deployed to
drill through porous, water-laden sedimentary layers (~3 km). MMW drilling can then be used to penetrate the
remaining 7-10 km of basement rock. The primary companies spearheading the R&D of MMW are Quaise Energy
(based out of the MIT Plasma Science and Fusion Centre), AltaRock, Advanced Research Projects Agency-Energy
(ARPA-E), Oak Ridge National Laboratory (ORNL), and Nabors Industries.
Quaise estimates that MMW could achieve an ROP of 3-5 m/hr with a 1 MW gyrotron for an 8-10” borehole
diameter, reaching 10 km in 100 days or less. MMW drilling is a relatively novel technique with a low TRL level and
has not yet been field-tested. As of February 2024, the deepest penetrated surface thus far is 2.4 m by 2 cm
diameter, performed with a low-powered beam. Research efforts to scale up experiments with a newly acquired
gyrotron with 10x the power output and 3x the frequency are underway.
Borehole construction in hard rock with MMW drilling could have several advantages over mechanical drilling
methods, namely: no performance limitations due to high downhole temperatures and pressures, minimal
downhole equipment (other than a metal waveguide), and automatic borehole reinforcement through rock
vitrification (Houde et al., 2021; Oglesby et al., 2014; Woskov & Cohn, 2009). There is no anticipated complication
at high temperatures, and vaporization may be more efficient in higher temperatures (Houde et al., 2021).
The MMW energy is supplied by a gyrotron, a vacuum electronic device that generates high-power
electromagnetic radiation at up to THz frequencies. The energy induces dielectric excitation at the rock’s ablation
front that is converted to thermal energy, raising the rock temperature past its vaporization point at over 3,000˚C.
To construct a wellbore, the MMW beam is transmitted through a metal waveguide made of a steel alloy with a
copper coating, which is continuously lowered into the well. To avoid damage to the waveguide from the high-
temperature ablation front, a stand-off distance is maintained between the rock and the waveguide (Figure 17).
The edges of the ablation front that are not fully vaporized melt and form a vitrified glass layer (Figure 18). This
dramatically reinforces the borehole walls and seals permeable zones from water influx, which may negate the
need for well cementation and casing once the well has reached hard, dry rock (Houde et al., 2021; Woskov,
2017).
Monitoring of downhole conditions may be communicated through remote diagnostic signals transmitted down
the waveguide. The ablation front temperature is monitored through radiometry, geological properties are
measured through exhaust spectroscopy, and the depth to the ablation front is monitored using
radar/reflectometry. This information may provide real-time recordings of ROP, downhole dysfunction, plume
41
depth, temperature, waveguide bends, lithology, fluids, and MWW wave mode changes (Oglesby et al., 2014).
Information may also be collected through the deployment of downhole sensors (Houderaque et al., 2021; Houde
et al., 2021).
Cuttings are removed by a circulating purge gas that conveys sub-mm scale particles from rock debris away from
the drilling front by an air drilling system at 100-5,000 psi. Noble gases are optimal as they maintain MMW
transmissibility at high pressures, with air used at shallower depths, followed by nitrogen, and then argon at 12-
15 km depths where nitrogen and air become supercritical. The purge gas also acts as a downhole cooling
mechanism of the waveguide (Houde et al., 2021).
42
Gaps and challenges for SHR
Experimental scale-up and ruggedization of the gyrotron apparatus are both an immediate focus for MMW-drilling
R&D to achieve TRL 5+ levels. Quaise recently acquired two 100 kW gyrotrons and is in the process of securing a
1 MW gyrotron by 2025. The manufacturing rate of gyrotrons globally is between 1-3 years for a single unit, but
this could accelerate once the demand increases and with the increased capacity of manufacturing facilities
(Woskov & Cohn, 2009). Partnerships with Nabors, AltaRock, ARPA-E, and ORNL will accelerate R&D.
Experimental bench tests at >374˚C have not yet been conducted, as the current research stand is at atmospheric
conditions. Acquiring a research stand capable of confining rock samples at supercritical conditions is a first-order
technology gap for MMW drilling. Along with bench tests, Quaise, with engineering support from Nabors and
ORNL, is in the process of designing a drilling rig that can accommodate the complex gyrotron system to conduct
in-situ field tests to achieve TRL 6 status. This rig will be deployed at a field location in Texas in 2024-2025, aiming
to penetrate 100-1,000 m (Houde et al., 2021). The rig will be designed to ensure that it can withstand field
conditions, namely rough roads to remote areas, and be operated by drilling personnel who are unfamiliar with
fusion technologies.
The dielectric absorptive properties of rock in the context of MMW drilling require further analysis through
modelling and experimentation. Different rock types will evaporate at different rates depending on their electrical
conductivity and absorptive properties (Woskov & Cohn, 2009; Woskov et al., 2014); therefore, MMW drilling
through heterogeneous formations may lead to an asymmetrical borehole (Figure 19). However, the uniformly
radial shape of the borehole is essential for the continued deployment of the waveguide downhole.
Granite and basalt are the likely rock types to be encountered within deep geothermal settings, yet even within
granite, variations in quartz content with low-absorptivity properties can induce heterogeneous vaporization
(Zhang et al., 2023). However, the absorptive properties of quartz and other minerals change at high temperatures
and in a molten state (Jerby et al., 2005). A greater understanding of these complex processes through numerical
and laboratory experimentation is required. These results will be essential in the MMW-drilling process and rig
design, as some form of adaptive MMW-energy input may be necessary to cope with the blind changes in lithology
downhole that are inherent in a natural setting.
43
Figure 19 Simulation results for multi-strata with heterogeneous absorptive properties, showing that the vaporization process does not
produce a uniformly radial well.
(Zhang et al., 2023)
The waveguide—while simply a metal tube—is a deceivingly complex component of MMW drilling. The waveguide
must retain a near-perfect vertical orientation to avoid rapid overheating from the MMW encountering the
waveguide. For example, a 30 m bent section of waveguide can heat to over 200˚C in 8 minutes from a 2 MW
beam, leading to 10 percent energy losses, damaging the waveguide (steel starts to degrade at 316˚C), and
potentially preventing its further deployment downhole (Oglesby et al., 2014). Tortuous or “corkscrew”
trajectories commonly seen in conventionally drilled wells may trigger this problem, which could be a considerable
challenge for MMW drilling because the proposed drilling program requires conventional drilling for the first 3-5
km of the well.
In the event of waveguide damage, the act of fishing an up to 10 km-long metal pipe from a vitrified well bore is
an unknown process and may lead to damage or complete loss of the well. The front of the waveguide must also
maintain its stand-off distance from the ablation front, or it will be at risk of overheating above its temperature
threshold. The collision of the waveguide on the active ablation front could only be prevented by a highly accurate,
real-time downhole monitoring system—one of the active areas of investigation by Quaise.
The wave guide must also be machined to very tight tolerances to maintain a continuous wave to the bottom of
the hole. But even if these tolerances can be achieved, the expansion and contraction of the waveguide at
different temperatures and pressures seems to be a major challenge. Typically, an EGS project requires two
directional wells to be drilled closely together either horizontally or at high angle, which requires significant
doglegs to be designed into the well. This type of drilling may be beyond the capabilities of a directional MMW
system.
44
Another challenge is hole stability in an underbalanced condition at great depths—a challenge shared by most
deep, hard rock drilling methods. A borehole from MMW drilling can theoretically be strengthened with insert
additives/pellets downhole to produce a glass/molten material that can help prevent hole collapse. Another
potential solution is to simply redrill the hole, after a collapse event occurs, repeatedly until the surrounding rock
relaxes itself to such a degree that there is not enough stress in the near-borehole field to displace rock back into
the hole. This may be possible since we do not have drill bits or downhole tools that would get stuck at the bottom
of the hole (but this is currently a speculative hypothesis).
Finally, while pore fluids within the rock do not pose a significant challenge to MMW drilling, highly fractured,
hydrous media or subterranean caverns will impede MMW transmission. Again, sealing pellets are a potential
solution—but this challenge may only be practically resolved through in-situ experiments.
MMW technology is in its infancy and needs a near-full-scale demonstration outside of a laboratory environment
to prove its potential. Practicalities in the deployment of the waveguide need to be addressed, including the
attenuation of the MMW due to diameter tolerances and expansion/contraction from variations in temperature,
pressure, and tensile load. Technology developers need a better understanding of the impact of MMW energy on
interbedded heterogeneous formations. They should also calculate and publish the cost per meter of MMW
systems.
The primary technology gaps for MMW drilling for SHR include:
Returning to the three core challenges that rock-destroying equipment must overcome to reach SHR domains
(drill to supercritical conditions >374˚C; drill to >5 km through hard rock where most SHR regimes are found; and
achieve rapid increases in ROP through hard rock, cutting back drilling time by >50 percent).
Conventional rotary and percussive methods have already overcome one or two of these challenges but no
method has conquered all three simultaneously. For example, Fervo and Utah FORGE have demonstrated leaps in
45
ROP generation through hard rock with PDC bits, however, they did not reach SHR temperatures or depths. IDDP
reached supercritical conditions with percussive and rotary drilling. At Stn-1, substantial ROP through hard rock
to depths of >5 km was achieved with percussive drilling, but this domain was relatively cold. Many hybrid-
conventional drilling methods have reached in-field experimental stages, ranging from a TRL status of 5-7. They
should be continually deployed at existing or future SHR projects to conduct necessary performance testing.
Primary challenges for advancing both conventional and hybrid-conventional methods to SHR capacity include:
• Limited access to SHR conditions in a controlled setting such as a laboratory bench test may be preventing
the iterative improvements required to optimize ROP generation;
• In-field experimentation of each method is needed to advance these technologies for hard, deep, hot
rock, to iteratively approach SHR domains;
• Most BHAs are designed with high-temperature ratings, up to 300˚C, although the remaining 75˚C will
require further innovation (except Strada Global, which penetrated supercritical domains for IDDP); and
• Coupling MWD and LWD sensors into these specialized drill bits may require further innovation and
collaboration between independent drilling firms.
Therefore, the primary gap for mature drilling technologies (conventional rotary and hybrid conventional) to drill
to SHR is scaling up experimentation with specialized high-temperature, hard-rock drill bits in SHR conditions or
analog EGS field sites.
Scaling up SHR testing will require a coordinated international effort, as SHR analogs are rare, and firms developing
high-temperature drilling technology may not reside in the same nations where SHR projects exist. However, deep
EGS and SHR projects are gaining momentum, and this research effort can be mobilized relatively quickly once
collaborative incentives are in place. Once these high-TRL technologies are actively experimenting in the field, it
is possible that one drilling method, or a combination of methods, may be fully optimized for drilling to SHR.
However, even with a fully optimal drilling system using a conventional or hybrid conventional approach, these
methods may reach a limit. Whether in drill string weight, ROP limitations, or excessive tripping time with
deepening wells, mechanical drilling technologies may not be the most effective method to reach SHR in the long
term.
Direct energy drilling methods are steadily advancing their R&D programs in the pursuit of SHR geothermal
projects. Should they reach commercial scalability, direct energy methods may prove more efficient than
mechanical drilling methods due to their minimal downhole equipment, reduced tripping or no need to trip for
bit replacement (increasing ROP), >7 km drilling capability, no adverse effects of high temperature on downhole
equipment (aside from the MMW waveguide that will experience thermal expansion/contraction), and reduced
energy consumption to drill relative to rotary or mechanical methods. These technologies are, however, still at
low TRL levels: TRL 3-4 for plasma and TRL 2-3 for MMW. To advance the frontier for direct energy methods to
reach SHR domains, both plasma and MMW drilling require:
46
3. Solutions to deploying downhole equipment (such as the waveguide for MMW drilling),
4. Modelling of the response of rock to direct energy exposure when the rock has heterogenous minerals or
properties,
5. Solutions to cleaning the drilling fluid from the borehole after completion of conventional drilling through
the first 2 km of soft overburden,
6. Methods to switch from the conventional drilling system to the direct energy drilling system, and
7. Training of operating personnel for direct energy drilling.
This kind of experimentation requires the creation (or adaptation) of laboratory facilities with SHR capacity to
bench test these technologies. Such a facility would likely require governmental funding and must be constructed
with compatibility with direct energy apparatus and rig design in mind. It must also be internationally accessible.
One candidate agency that could receive this funding is Nabors Industries, as they are partners with both GA
Drilling and Quaise Energy and are actively working with them on rig design. Nabors also has drill stands at
atmospheric conditions, which could possibly be adapted to simulate SHR temperatures and pressures.
In summary, the three drilling categories are at the following frontiers for SHR drilling:
1. Conventional rotary drilling is in the lead but may reach a performance limit,
2. Hybrid conventional drilling may prove superior but requires further experimentation, and
3. Direct energy drilling has long-term potential after significant experimentation and field testing.
47
5. High-temperature downhole tools: Technology frontier and gap
analysis
In geothermal drilling, “downhole tools” refer to the various instruments and devices used in the wellbore to
perform operations, gather data, and manage the drilling process. Key components of downhole tools include
components of the BHA, which comprises stabilizers, reamers, shock tools, mud motors, rotary steerable systems
(RSS), Measurement While Drilling (MWD) tools, and Logging While Drilling (LWD) tools. Additionally, downhole
sensors are essential for monitoring and gathering data during the drilling process. The ability for these tools to
operate in high-temperature, high-pressure environments is vital to the success of a drilling project.
To achieve a successful well in SHR, downhole tools must withstand high temperatures, and/or the borehole must
be cooled to their maximum operating temperature threshold. The temperature limitations of downhole tools
must be considered in combination with the technologies and processes discussed in Section 4. If the wellbore
can be cooled to below 150˚C, then the system is a TRL 9. However, running 150˚C equipment in the hole would
require a full string of insulated drill pipe (IDP). To run a full string of IDP requires a significantly larger rig, a higher-
pressure surface system, and greater hoisting capacity. The increase in rig size and the additional cost of the IDP
may incur costs that make the SHR project uneconomic.
The most cost-effective solution is a combination of high-temperature downhole tools and technologies from
Section 6 (temperature management equipment). Even if the borehole is cooled to below 150˚C (normal O&G
electronic tool operating temperature), a significant temperature safety margin must be maintained for
connections, loss of circulation, surface equipment failures, and “staging in.” Staging in is the process of deploying
the BHA and drill pipe into the hole and circulating fluid to cool the borehole. This process incurs invisible lost time
(ILT) and adds significant cost to an operation—a phenomenon that was observed at FORGE when drilling the
deepest section of the well.
Pressure limitations of most high-temperature tools are in the range of 25,000-30,000 psi, which is within the
range of SHR wells. Due to low mud weight, SHR geothermal wells will likely have bottom hole pressures under
20,000 psi. The main limitation of these tools is the maximum temperature rating of the elastomers in the pressure
seals. Viton rubber seals have a max temperature of 300⁰C and Tetrafluoroethylene O-Rings 232⁰C.
High-temperature tool limitations are currently around 175⁰C to 200⁰C, and several companies are striving for
temperatures >210⁰C. Recent technology demonstrations at the Lincoln Laboratory at MIT showed that the
temperature limitation for electronic components could be cost-effectively raised to 300⁰C.
One major challenge for high-temperature downhole tools is power supply. The maximum operating temperature
of downhole battery packs is 200⁰C. This limitation can be overcome with a downhole turbine that provides power
to the tool. High-temperature electric motors for downhole turbines with non-organic insulation are achieving
higher temperature ratings (also demonstrated at the Lincoln Laboratory). A turbine can cause problems for
measurements that require a continuous supply of power.
48
5.1 High-temperature tool manufacturers
Hephae is a startup operated by former employees of major service companies and downhole tool manufacturers.
Their near-term focus is on 210oC tools, with aspirations for increasingly higher temperatures. Hephae is designing
and building an MWD called PandoraX and a rotary steerable system (RSS) that is modularized and can be linked
to their PandoraX platform (Figure 20).
The use of high-temperature (>210⁰C) RSSs will be pivotal to drilling SHR, as developers will need greater precision
and better real-time data from MWD tools. For example, adding a mud motor to the bottom of the drilling system
often creates RPMs that are too high for optimal depth of cut. The pressure losses through the motor indicate
that flow rates need to be lowered, in turn reducing the amount of fluid available for cooling. Further, rubber
components in mud motors lower the maximum operating temperature and reliability of the combined drilling
system. According to John Clegg at Hephae (Interview March 2024), the high-temperature RSS allows full
directional capability with continuous rotation, which improves hole cleaning and reduces the risk of the BHA
getting stuck downhole.
In Section 6 (temperature management equipment) we conclude that even if an IDP was available for a high-
temperature project, there would be significant commercial and technological advantages to running the IDP in
conjunction with >210⁰C downhole tools, which could reduce the amount of IDP by around 50 percent, lower the
string weight, and improve hydraulics. The estimated premium of IDP over regular drill pipe is roughly $6,000 (all
costs in this paper are USD) per joint. If a developer can run 210-250⁰C tools and reduce the amount of IDP by 50
percent in a 25,000 ft well, savings on capital costs would be around $2.5 million per well. Another major cost to
SHR geothermal is the rig time spent circulating and cooling the wellbore while running back to bottom. Combining
high-temperature tools with a 50 percent insulated string and low heat coefficient coated drill pipe could reduce
trip costs by USD $50,000-$100,000 a trip.
NOV have developed a range of high temperature tools, including the Tolteq™ iSeries and the Tolteq™ Hellfire.
These tools go through a rigorous temperature qualification process, cycling the temperature on the tool from
25°C to 175°C for 1,000 hrs. The Tolteq™ iSeries platform is the leading probe-based, mud pulse MWD available
in the independent directional drilling market. As a modular platform based on the legacy Tensor design, the
49
iSeries platform provides significant flexibility to configure and deploy the tools in a wide variety of applications.
According to Anthony Dukowski and Paul Neil of NOV (Interview Feb. 2024), the NOV Tolteq™ tool range is well-
suited for geothermal operations in remote locations. One of the advantages of running probe-based tools into
the well is that if temperatures start to become too hot or the string gets stuck, a wireline can be run into the hole
to retrieve the tools, reducing lost-in-hole costs.
The Weatherford Heatwave tools provide a 200°C-rated “triple combo” LWD/MWD service. These tools—used in
combination with IDP, mud coolers, and low heat coefficient coatings—can operate in SHR formations up to
400°C.
Gunnar Energy Services is actively pursuing advancements in high-heat drilling, particularly for active and passive
magnetic ranging technologies. The technology has been proven to be effective up to 220°C for Steam Assisted
Gravity Drainage wells but is restricted by the heat tolerance of MWD sensors. The potential to adapt these
ranging methods to operate at SHR temperatures is a significant technological challenge. To achieve this, Gunnar
plans to develop pipe and active ranging solenoid components with ceramic coatings designed to withstand
extreme heat. While the theoretical foundations for these adaptations are established, they require
comprehensive testing to understand how supercritical heat affects magnetic signatures in controlled laboratory
environments and real-world field trials.
According to Andrew Penman and Matt Holdeman Halliburton (Interview, February 2024), Halliburton has done
extensive development on high-temperature tools, with one line rated at 175°C (Solar®) and another at 200°C
(Quasar®). Their maximum operating temperature and pressure is summarized in Table 1. The Quasar® line of
tools have a high temperature rating but were designed for high-temperature oil and gas wells. The diameter of
the Quasar® tools is not suitable for SHR geothermal, as they are too slim for 8 ½” hole sections.
50
Table 1 Halliburton’s high-temperature tools.
(Halliburton, 2024)
Directional wells require both measurement tools and a drive system. Halliburton is developing motors with high-
temperature elastomers rated up to 188°C, and an RSS rated up to 175°C and 30,000 psi. For the well-on-well
connections necessary for drilling a closed-loop SHR well, Halliburton’s rotary magnetic tools (rated to 175°C)
could be combined with other temperature management equipment to work at 400°C.
Baker Hughes has designed a drilling system to cope with 300˚C temperatures at 10 km depths in hard rock for a
minimum of 50 on-bottom hours specifically for SHR. The system includes an MWD, drill bit, steerable drilling
motor, and drilling fluid for an 8 ½” borehole (see Stefánsson et al. 2018 for the complete review of the system,
including its performance at IDDP-2).
Deep directional drilling requires positive displacement motors (PDM), that include a “drive sub” that connects to
the bit, an adjustable kick-off that enables bending in the drilling system, and a power section (Figure 21). A point
of weakness in conventional PDMs is the elastomer sealing that line the stators, which are rated to 190˚C.
Baker Hughes has implemented full metal stators in their high-temperature drilling system, referred to as the
metal-to-metal (M2M) motors, that are composed solely of steel alloys with a mud-lubricated assembly and are
rated up to 300˚C. The rotor and stator are elastomer-free with a wear-protection coating that protects the metal-
51
to-metal contact points. One drawback is the potential for fluid leakage between the rotor and stator without the
elastomer sealing; however, a fully sealed chamber could be engineered with further R&D. In-field and laboratory
experiments proved that high torque outputs can be maintained at temperatures over 250˚C for over 50 hours,
with an anticipated drop in rotational speed from 80-50 RPM throughout the drilling run (Stefánsson et al. 2018).
Downhole MWD and LWD instruments contain electronics that cannot withstand SHR temperatures. Rather than
developing costly electronic tools with higher-temperature ratings, Baker Hughes developed a cooling system for
their BHA to maintain internal temperatures at 175˚C, even under 300˚C conditions. According to John
Macpherson (Interview, 2024), Baker Hughes’ cooling system capitalizes on the evaporative properties of water
and isolates the temperature of the thermos flask containing the electronics through a control valve.
MB Century (350°C)
While drilling and completing an SHR well, it may be necessary to run tools into the wellbore to log the
temperature, casing quality, and cement quality. Logs can also be run regularly to understand how the well is
changing due to thermal cycling and corrosion. According to Greg Thompson (Interview, 2024) MB Century’s tools
have specifications very close to SHR conditions. Logs can also be run regularly to understand how the well is
changing due to thermal cycling and corrosion.
Schlumberger has developed an MWD and an RSS for 200°C. The TeleScope ICE (MWD) and the PowerDrive ICE
(RSS) tools have both been tested for 2,000 hours at 200°C. Both tools have integrated ceramic electronics (ICE)
and multichip modules designed specifically for high-temperature environments. These tools are collar-based, and
unlike the probe-based tools, they are not retrievable if the wellbore temperatures get too high or the assembly
gets stuck in the hole.
The temperature limitation for most downhole tools is within 175-200°C, with some tool manufacturers reaching
220°C in their pursuit of SHR. The most cost-effective and market-ready approach is to deploy existing high-
temperature tools in a well cooled to the operable temperature with the equipment discussed in Section 6. The
combination of these methods is at TRL 9. Technology companies must continue to improve high-temperature-
resistant electronics with non-organic materials (such as all-metal seals) and develop high-temperature-resistant
capsules for electronics. Further research into “one run” tools that are deployed once before expiring from high
temperatures may provide a cost-effective (but potentially wasteful) alternative to engineering SHR-capable tools
or deploying temperature management equipment in combination with existing high-temperature tools.
52
5.2 High-temperature motors
Several publications and papers (e.g., Stefánsson et al., 2021; Epplin, 2015) discuss the use of mud motors with
metal-on-metal power sections. These motors all appear to have a temperature rating of around 300°C. Based on
discussions with various groups deploying these motors, the authors conclude that challenges remain in machining
the steel parts within the required tight tolerances—and, even when those tolerances are met, the motors can be
susceptible to erosion by particles in the drilling fluid.
High-temperature motor technology is at TRL 6 and requires further investment, field testing, and ruggedization
to reach TRL 9. If a drill string is designed to cool the well to 175°C (the current limit of electronic tools), it may
prove unnecessary to develop motors that reach 300°C. In that case, operators can opt for high-temperature,
even-wall rubber motors rated to 190°C.
5.3 Conclusions
No high-temperature sensors are rated for SHR. The practicality of building tools to survive in well temperatures
of >374°C without the use of any temperature management systems would require significant scientific
breakthroughs. The National Renewable Energy Lab and NOV spent a year investigating 374°C electronics to
determine if it was possible to design tool components using packaged aluminum gallium nitride/gallium
nitride (high electron mobility transistors are used in high-frequency and high-power switching applications). This
study showed that, even if chipsets could be built with this compound, the batteries and other components
remained a limitation (based on the experience of the lead author at NOV).
The consensus of equipment manufacturers is that electronics will only withstand SHR conditions with a
combination of the highest-temperature-rated tools and a cost-effective temperature management system. As
described in the next section, managing the temperature of the borehole to below 175°C is achievable today, but
temperature management equipment requires further testing and validation in SHR conditions. Therefore, the
TRL of high-temperature sensors in SHR conditions is TRL 8 when used with other cooling methods, TRL 7 if
encapsulated, and TRL 3 for stand-alone tools.
53
6. Temperature management equipment: Technology frontier and
gap analysis
When drilling fluid gets too hot, its ability to carry the drilled cuttings deteriorates, bit wear increases, downhole
tools are destroyed, and the risk of drill pipe and casing corrosion increases by orders of magnitude. Overheated
drilling fluids also prevent the use of downhole motors and electronic instrumentation and can affect borehole
stability and well control. Each of these consequences can be countered by individual technology developments,
but all can be solved (or greatly mitigated) by simply cooling the downhole environment.
Temperature management can be achieved with a combination of equipment, including low-heat coefficient
coatings, mud coolers, and (most significantly) insulated drill pipe (IDP). Current drilling technologies can be used
in 400˚C rock; however, to do this economically, a combination of technologies must be deployed. For example,
the combination of high-temperature tools with low-heat coefficient coatings and mud coolers decreases the
amount of IDP needed, lowering daily operating costs. IDP generally has a smaller inner diameter and is
significantly heavier than traditional drill pipe. The rig would thus require higher-pressure-rated pumps and
standpipe, a stronger derrick, a stronger substructure, and draw works with a higher lifting capacity. This increased
rig capability would enable the SHR well to be drilled but at a significantly higher cost.
An alternative method to managing borehole temperature is to drill with CO2 in its various states: gaseous, liquid,
and supercritical fluid (sCO2). sCO2 has a similar density to a liquid and a viscosity comparable to a gas. The sCO2
can be used as a drilling fluid to keep the drilling tools cool, and its density could power downhole mud motors
and RSS. The downhole pressure and temperature domain would hold the CO2 in a supercritical state on the
journey down the drill pipe or coiled tubing, and it would convert to vapour when exiting the drill bit nozzles. With
the CO2 in a vapourized state, the borehole annulus will be in an underbalanced condition, therefore less likely to
lose circulation (a common problem in geothermal wells, especially wells in basement rock with a high fracture
density).
A core challenge with using CO2 as a drilling fluid is its availability in a concentrated supply. One O&G operator
that drilled several CO2 wells in Louisiana managed to source CO2 at a volume and price required for enhanced oil
recovery, but these resources are in very limited supply. Significant advances in carbon capture and a reduction
in cost are required to scale up CO2 drilling.
Thermal and mechanical analyses on insulated drill pipe (IDP) were first conducted at the Sandia National
Laboratories in the 1980s but the concept remained relatively dormant for nearly 20 years. In 2000, Drill Cool
Systems built three joints of prototype IDP, conducted preliminary tests to evaluate the effective thermal
conductivity of the pipe, and ran these joints in field operations. In a collaboration with the Geothermal Drilling
Organization, they constructed a complete string of pipe (5500’) and tested it in Imperial Valley in California (see
Champness, 2000).
54
There are three basic approaches to IDP. The advantages and disadvantages of these approaches were defined
well by the Sandia team in 2000 and are summarized in Table 2 below.
Several recent studies have investigated the effectiveness and potential of IDP in SHR applications. (Ajima, K.,
Naganawa, S., 2022; Pink et al., 2023). The study by Ajima and Naganawa modelled the use of IDP in a 4,000 m
well with a max temperature of 600˚C. They used the base pipe design from Sandia and Drill Cool and GEOTEMP2
software. The model showed that a full string of IDP could deliver sub-175˚C fluid at 4,000 m under supercritical
conditions.
When they investigated the effects of running a half string of IDP (flowing at 2,000 L/min), the temperatures were
209.6˚C (upper half of IDP) and 233˚C (lower half of IDP) at 4,000 m. Both temperatures are higher than the normal
operating temperatures of electronic tools. However, there was a significant effect when the flow rate was
increased from 2,000 L/min to 3,000 L/min. At 3,000 L/min, the bottom hole temperatures were lowered to 132˚C
(upper half of IDP) and 159˚C (lower half of IDP). These results are depicted in Figure 22 and summarized in Table
3.
55
1500 (L/min) 2000 (L/min) 2500 (L/min) 3000 (L/min)
Upper Half (˚C) 280.2 209.6 163.0 132.0
Lower Half (˚C) 286.9 233.6 191.6 159.8
Table 3 Bottom hole temperature for different mud circulation flow rates in a partial (1/2) IDP application (adapted from Ajima et al.,
2022).
Figure 22 Temperature profiles when applying IDPs along (Left) the upper half of the wellbore and (Right) the lower half of the wellbore
with drilling mud.
(Ajima et al., GRC 2022)
In 2022, NOV investigated the limitations of the original Sandia and Drill Cool IDP designs. They focused on two
areas: (1) applying low heat coefficient coatings to the internal diameter (ID) of normal drill pipe and (2) designing,
modeling, and building a string of IDP that did not have the ID limitations of the Sandia-designed pipe (3”).
The NOV team presented their progress on IDP designs at the Society of Petroleum Engineers conference in
Stavanger in 2023. Their IDP consisted of a unique thread form in the box end of the connection (Figure 23) that
enabled a 4” rubber-coated inner tube to be pressed into the drill pipe body. The ID of the NOV IDP is 3.69”,
significantly larger than the original 3.0” design. The NOV team also reported that the thermal conductivity of the
pipe could be lowered significantly by applying a low heat coefficient coating to the inside of the tool joints,
removing some steel from the inside of the tool joints and replacing it with Inconel or stainless steel, and using
range 3 pipe (therefore 30 percent fewer tool joints in the string). Lastly, they highlighted the effect of heat
capacity of mud versus water. If mud is replaced with water (which has a lower specific heat capacity), the bottom
hole temperatures are higher.
56
Figure 23 IDP design from NOV that allows an inner diameter of 3.69” inside a 5 7/8” drill pipe tube.
(Pink et al., 2023)
In the IDP design above, the whole of the IDP string is coated with a low heat coefficient coating provided by a
division of NOV called Tuboscope. It is additionally important to consider the effect of heat capacity on mud and
water. In its current form, IDP is heavy, relatively expensive (roughly 2-3x the cost of regular drill pipe), and has
high pressure losses—placing higher demands on pump pressure capability, the strength of the derrick and the
substructure, and the lifting capability of the draw works. However, running IDP and cooling the fluid to maintain
surface temperatures below 100˚C would have a significant positive impact on the cost of other rig equipment
and would eliminate the need for high-temperature elastomers.
Looking at the drill string (not including mud coolers and mud properties), it is the opinion of the authors that the
currently available IDP technology combined with coated regular drill pipe can deliver sub-175˚C fluid to the
bottom of a 7,000-8,000 m well in SHR conditions. Operators should look for the lowest-cost, highest-value
combination of coated pipe, IDP, and high-temperature downhole tools (along with a surface rig that delivers the
optimum flow rate with sufficient hoisting capacity). If the IDP drill string is too heavy, higher -temperature-rated
tools could be run and/or surface mud coolers could be added to the system.
Recent modeling shows that there are no major technology gaps to managing the temperature of a 400˚C SHR
well below 175˚C while drilling (Pink et al., 2023). However, the technology must be manufactured, and a full-
scale pilot project must be drilled at SHR conditions to validate the technology and ensure that it is field-hardened
for the downhole conditions. The complete drilling system must be modelled prior to in-field testing to ensure the
most cost-effective solution is selected. Drill Cool has a market ready product (TRL 8) and NOV has a design built
on a platform of known technology but requires a project and customer (TRL 6).
Titanium is well-established in aerospace, marine, and defense industries for its outstanding performance in
environments above 500˚C. The metal's low thermal conductivity insulates drilling fluids within the pipe from
extreme external heat to maintain lower temperatures within the pipe and increase drilling efficiency. When
paired with low heat coefficient coatings, titanium drill pipe can further reduce thermal transfer, boosting drilling
rates by promoting temperature-induced fracturing at the bit face.
57
Figure 24 Images of titanium drill pipe from ALTISS Technologies.
(ALTISS Technologies, 2024)
Titanium’s excellent strength-to-weight ratio enables the production of lighter yet stronger drill pipe and well
tubulars. This reduction in weight improves the depth capabilities of drilling rigs, broadens rig selection and
options to control rig move costs, and reduces the footprint of drilling pads. Further, titanium's superb corrosion
resistance and thermal stability at temperatures up to 500˚C ensures the long-term durability of tubulars and
downhole equipment in harsh conditions with far longer life cycles than stainless chrome and nickel alloys. These
longer life cycles significantly decrease the need for maintenance and replacement in severe downhole conditions.
Titanium also exhibits superior fatigue resistance and dimensional stability, crucial for maintaining structural
integrity during severe thermal cycling. This property makes it particularly suitable for scenarios with frequent
temperature fluctuations, effectively preventing fatigue cracks and deformations that could lead to system
failures.
58
ALTISS Technologies is at the forefront of titanium drill pipe designed for hostile geothermal conditions, including
SHR conditions. Their patented titanium drill pipe (Figure 24) dampens vibrations to enhance mechanical energy
efficiency, leading to higher penetration rates and extending the life of the BHA and drill string. In addition to drill
pipe, ALTISS is applying titanium to other components, including drilling and completion systems components,
casing, and tubing. This underscores titanium’s growing importance and ongoing improvements in high-
performance applications.
The cost of titanium drill pipe is the biggest challenge for scaling the technology. Testing alternative alloys, metals
and coating solutions should be prioritized for SHR development. If costs can be brought down, titanium drill pipe
would be at TRL 9.
Drilling fluid maintains downhole drilling temperatures and mud stability, suspends cuttings, and protects the
wellbore. These functions are essential in geothermal domains where the borehole is at risk of thermal and
structural instability, circulation losses, and high frictional resistance (Petty et al. 2020; Song et al. 2023). The fluid
requires enough viscosity to form a “wall cake” along the borehole surface to prevent fluid losses, typically
achieved with biopolymers in conventional drilling (which are only rated to <149˚C) (Petty et al. 2020). Most
geothermal muds are a mixture of fresh water and bentonite clay, with polymers added to improve mud viscosity
and filtration properties. Aerated muds may also reduce circulation losses in geothermal settings (Song et al.
2023).
From a purely temperature management point of view, running a mud with a higher specific heat capacity than
water would be beneficial; however, data from FORGE has shown that drilling with water with little or no additives
has a significant positive effect on ROP. The volatility and chemistry of drilling mud under high pressure and
temperature conditions are key factors. High temperatures may induce the flocculation (i.e., clustering) of
bentonite dispersion and degrade mud chemistry and viscosity, leading to filtration loss, reduced cutting
suspension and transport, and, ultimately, compromised borehole stability. The inclination and depths of
geothermal wells may increase frictional resistance, further compromising cutting transport. Interbedded
formations found at geothermal sites may also result in heterogeneous pressure distribution downhole and high
brine content may impact mud density and chemistry (Song et al. 2023).
To compensate for these risks, geothermal projects use additives to improve well stability, such as sulfonated
asphalt powder, sulfonated lignite, and sulfonated phenolic resin. For example, MAGCOBAR’s high-temperature
drilling mud, DURATHERM (a water-based mixture of clays, resins, and lignite) was used to drill a 3,200 m deep,
353˚C geothermal well at the Salton Sea. The German Continental Deep Drilling Programme (KTB) also developed
a silicate compound called Dehydrill HT with 280˚C rheological stability that can enhance drilling mud.
59
Synthetic polymers <204˚C can be added to the drilling mud, as well as clay-based drilling fluids such as
Halliburton’s ilmenite-based MicrodenseTM mud system used at Larderello for the project DESCRAMBLE (Petty et
al. 2020). The mud for the IDDP-2 well was rated up to 300˚C and remained stable after 50 hours of use (Figure
25) (Stefánsson et al. 2018). This mud was water-based, with a bentonite stabilizer and low molecular weight
copolymer additives and vinyl-sulfonated copolymers (Petty et al. 2020). Additional lubricants were required to
maintain the elastomer-free MWD deployed at this site (Stefánsson et al. 2018). Finally, Baker Hughes has
developed 260˚C rated viscosifiers, thinners, deflocculants, and filtration reducers that can be added to bentonite
slurries (Song et al. 2023).
Figure 25 Drilling fluid shear strength [lbf/100ft2], plastic viscosity (PV) [cp], yield point (YP) [lbf/100ft2] after exposure to 56˚C and 300˚C
(Stefánsson et al., 2018)
In geothermal well construction, maintaining wellbore stability requires drilling fluid with high thermal stability
and better rheological properties. Commonly used water-based muds and additives experience common
problems, such as lost circulation, wellbore cleaning, hydraulics, and fluid stability (Mohamed et al., 2021). Lost
circulation is the most frequently encountered, time-consuming, and costly problem faced during hydrothermal
well drilling. As the geothermal resources are in the fractures and fissures, lost circulation is expected as the target
depth is reached or whenever fractures are encountered.
The first step to combat this issue is to stop drilling and circulate the lost circulation materials. If the issue is still
not resolved, the next step is plugback cementing. The time required to resolve lost circulation issues is counted
as Non-Productive Time (NPT). In wells analyzed by Visser et al. (2018), the NPT due to lost circulation was as high
as 197 hours (8 days). Lost circulation accounts for 85 percent of the total NPT for some geothermal wells. The
cost of the project is highly affected by the time required to mitigate the lost circulation issue and the method and
materials used for mitigation.
Lost circulation risks can be significantly reduced by not drilling fractured basement rock and instead focusing on
SHR in dry conditions and then creating a fracture network after the well has been drilled. Hydrothermal wells are
designed to intersect natural fractures. By switching to an enhanced geothermal system (EGS) or closed-loop
geothermal (CLG) design, the well can avoid natural fractures and reduce the risk of lost circulation.
60
Gaps and challenges for SHR
Water, at the correct fluid velocity, is sufficient to clean the hole if developers maintain the borehole inclination
below 20 degrees. Above 20 degrees inclination, research is needed in high-temperature additives that would
provide adequate rheology to clean the hole, especially if operators begin drilling horizontal SHR wells. R&D in
high-temperature fluids could be conducted by national labs and universities in partnership with a commercial
entity who can test the technology. Advanced testing equipment that can perform a full set of mud tests at 400˚C
probably does not exist and needs to be designed and built. Drilling fluids in SHR conditions below 20 degrees
inclination have a TRL of 8 or 9, and more complex high-temperature mud systems for wells above 20 degrees
inclination are at a TRL of 4 or 5.
To use of supercritical CO2 (sCO2) as a drilling fluid—which is distinct from using sCO2 as a geothermal working
fluid—several modifications are required to make the system compatible with a conventional rig. The sCO2 would
be injected at the surface into the drill pipe and the supercritical state would be maintained by controlling the
flow rate and the operating pressure of the system. Such a system would be comparable to a conventionally
managed pressure drilling system that is used on operations where the reservoir is under-pressured and cannot
be drilled with normal mud weights. The pressure drop of the system would be dictated by the pressure drop of
the drill string, the downhole tools, and (primarily) the drill bit. The sCO2 would have sufficient density and provide
sufficient hydraulic power to drive the downhole tools and provide a medium for the tools to transmit a signal
back to the surface. As the sCO2 exits the nozzles in the drill bit, it changes state from a dense sCO2 phase to a
gaseous phase, causing a significant drop in pressure and temperature at the drill bit and rock interface. This drop
in temperature and pressure creates thermal impact on the formation, especially if the differential between the
formation temperature and the drilling fluid temperature is >100˚C (Phuoc et al., 2020). This thermal impact
weakens the crystal structure at the basement rock and bit interface, improving drilling efficiency.
As the sCO2 changes phase to a gas, it provides adequate annular velocity to clean the cuttings out of the hole.
Gaseous CO2 would also insulate the sCO2 in the drill pipe from the formation temperature, provide an
underbalanced condition in the annulus, and minimize the risk of lost circulation should there be any permeability
or large natural fractures.
The primary benefits of using sCO2 as a drilling fluid are: the ability to cool the drilling system so conventional or
high-temperature rated tools could be run in combination with the sCO2 system, the lower annular pressures
which significantly reduce the risk of lost circulation in naturally fractured formations, and potential increases in
ROP from the thermal impact of hitting SHR with sCO2. The primary disadvantages and complications of running
a sCO2 system are: sourcing a sufficient amount of sCO2, the environmental impact of potential leaks, and the
need for crewmembers with experience drilling with sCO2.
61
Gaps and challenges for SHR
All the components of a sCO2 system exist, but the complete system needs to be engineered and fully integrated
into a package that includes all the mechanical hardware, a control system, a model, and trained experts to run
it—and then that package must be tested in the field. With the right level of investment, this package could be
realistically developed and tested in 2-3 years. The TRL level for CO2-based fluid is 3, but this could quickly move
to a TRL 7. However, CO2 production needs to increase to scale up this technology.
An engineered system that optimizes the fluid dynamics may further enhance the cooling mechanisms employed
in an SHR well. Fluid dynamics optimization is achieved with an accurate well design, particularly regarding hole
size and drill pipe to deliver the optimum volume of cooling fluid to the drill bit. If a state of laminar flow can be
created in the annulus, then significantly less heat would be transferred from the hot formation to the drilling
fluid. The laminar flow effectively creates a layer of fluid close to the borehole wall that insulates the fluid. If
turbulent flow occurs, this layer would be destroyed, and the fluid would warm up significantly.
The well must be modelled prior to design phase and that model should be run in real-time to ensure that the
annulus is in laminar flow whilst drilling. With particular designs and flow rates, laminar flow may not be possible.
This technique would typically be used with other cooling methods, such as insulated pipe and mud coolers.
The TRL status for fluid dynamics optimization is 9, as the technology and physics are fully understood today. Fluid
dynamics optimization is a complementary process that enhances the new well-construction technologies being
developed.
Before the development of low heat coefficient coatings, previous coatings had an average thermal conductivity
of 0.8360 W/mK, which is much lower than carbon steel (45 W/mK). Through iterative formulation adjustments,
NOV, in collaboration with Eavor, has developed a coating formulation with an average conductivity of just 0.1808
W/mK, well below the conventional coating value and the operator’s target (Figure 26 and 27).
NOV and Eavor’s TK™-18TC is a low thermal conductivity coating designed with better insulative properties while
still retaining the same downhole performance (hydraulic performance/surface roughness, abrasion resistance,
chemical resistance, impact resistance, and flexibility). The coating has a lower thermal conductivity, allowing the
end user to better maintain the temperatures of the fluid within the pipe, and has an applied thickness of just 20-
30 mils (thousandths of an inch). As with all internal coatings, TK™-18TC is also designed to extend the life of the
tubular through corrosion and deposit mitigation.
62
Figure 26 Low heat coefficient coating for drill pipe.
(NOV, 2023)
Figure 27 Thermal conductivity comparison between TK34XT and the low heat coefficient coating TK™-340TC/18TC.
(NOV, 2023)
63
Low heat coefficient coatings can be used independently to moderately reduce borehole temperature or used
with insulated drill pipe and mud coolers to deliver much larger temperature reductions. Models have
demonstrated that the temperature could be maintained below 150°C (Pink et al., 2023, SPE Stavanger).
Another company, Tuboscope, is also developing low heat coefficient coatings with even lower thermal
conductivities, with the aim of helping future geothermal operators drill into deeper reservoirs with temperatures
exceeding 300°C. Such innovations will also help operators in shallower, lower-temperature reservoirs minimize
heat losses in the pipe to support applications such as district heating, low-enthalpy electricity generation,
greenhouses, and hydroponics farms.
Low heat coefficient coatings are used commercially in geothermal wells and other deep hot wells. When
combined with mud coolers, insulated drill pipe, and other temperature management equipment, low heat
coefficient coatings are at a TRL of 9 for SHR wells. Without these added components that improve insulation and
cooling, they are at a TRL of 6. Continued research into low heat coefficient coatings and the potential use of
nanoparticles and other insulative materials may further lower the conductivity. This research would be carried
out most effectively through collaborations between national labs and commercial test partners. The low heat
coefficient coatings may also be considered for insulative purposes on other tubulars in the wellbore.
Surface mud coolers are a relatively cheap solution to reducing borehole circulating temperatures. Mud coolers
are effective in relatively shallow superhot geothermal wells, but if used with conventional drill pipe, the
effectiveness diminishes with depth (Khaled et al., 2023).
Figure 28 shows that mud coolers effectively reduce the bottom hole circulation temperature (BHCT) of shallower
wells (see Well A, 2,657 m). The BHCT in Well A decreases from 128 to 109˚C (a 15 percent temperature reduction)
when the surface’s mud temperature decreases from 70 to 30˚C. However, for deep geothermal wells (Well B), a
mud cooler does not significantly reduce the BHCT while drilling if used without another insulative technology.
The BHCT in Well B (2,617 m) is reduced only by 1 percent (from 179 to 177˚C) when the surface temperature of
mud is decreased from 70 to 30˚C.
As the well gets deeper (with constant circulation rate), the total travelling time and distance of the drilling fluid
to reach the bottom hole increases. When a mud cooler is used, the temperature difference between the
formation and the drilling fluid is increased, which increases the heat flux per unit wellbore length from the
formation to the drilling fluid. The temperature rise of the drilling fluid when travelling to the bottom is
proportional to the total wellbore length (L), and the average temperature difference between the drilling fluid
and the formation (ΔT).
64
Figure 28 Impact of mud surface temperature by a mud cooler on BHCT of wells A & B.
Note the reduced effect of the mud coolers action for well B (Khaled et al., 2023).
Both NOV and Drill Cool manufacture mud coolers that could be utilized for cooling superhot geothermal systems
while drilling. NOV manufactures a range of land mud coolers under the brand name Tundra™ Max, which operate
with a flow rate of 350 to 500 gpm. If higher flow rates are required, units must be used in parallel or in series to
lower the outlet flow temperature for a given flow rate. The Freshwater Geo‑Cooler™ from Drill Cool incorporates
a closed-circuit cooling tower with a DNV-certified offshore ISO container.
Although results show that mud coolers are not an effective heat management strategy for SHR wells, they still
offer benefits when used during drilling by maintaining low surface mud temperature. Lower surface mud
temperatures increase crew safety and helps avoid mud pump failure caused by high-temperature mud. In
addition, the effectiveness of mud coolers is increased when combined with IDP. Mud coolers that can be
deployed for SHR projects exist today (and are therefore at TRL 9).
6.3 Conclusions
An SHR well must be cooled to operable temperatures during drilling operations, as high temperatures pose a risk
to bit wear (Section 4) and downhole monitoring systems (Section 5). Deploying an optimal combination of existing
and improved temperature management equipment—namely, mud coolers, IDP, low heat coefficient coatings,
and drilling fluid—is likely needed to drill into SHR. Finding the optimal combination of these systems will require
further high-temperature bench tests and in-field experimentation.
To overcome the mounting cost and weight incurred by excessive lengths of IDP, alternative technologies such as
sCO2 drilling and titanium drill pipe should be investigated. The major obstacle to sCO2 drilling is the supply of CO2,
while the major obstacle to titanium drill pipe is cost. Scaling up SHR drilling may improve the availability and
economics of these technologies.
65
7. Corrosion inhibition: Technology frontier and gap analysis
Corrosion inhibition is a major concern in geothermal projects, particularly for conventional hydrothermal systems
where deep formation fluids are produced. For example, the brines from wells in the Salton Sea have a high Na-
Ca-K-Cl type (approximate salinity 185,000 ppm Cl) with exceptionally high potassium, as well as the highest
content of minor alkali elements known for natural water. Fe, Mn, Zn, Pb, Cu, Ag, and other metals are also present
in exceptionally high concentrations. However, other waters produced from geothermal wells around the world
are of near drinking-water quality and are noncorrosive.
Corrosion inhibition is largely beyond the scope of this paper as this concern is primarily for production companies
rather than drilling operators. The drillers must, however, consider the impact of corrosion on drilling equipment,
downhole tools, casing metallurgy, and the design of the cement. All the corrosive processes associated with
drilling geothermal wells would be exaggerated under SHR conditions, which would accelerate the chemical
processes.
Drill pipe can be protected with coatings. For example, corrosion-resistant coatings have been incorporated into
the low heat coefficient coatings produced by NOV (Figure 29). Titanium drill pipe is very resistant to corrosion but
is expensive compared to steel pipe.
The chemistry of the formation fluids and the presence of chlorides, carbon dioxide, and hydrogen sulphide also
have an impact on the well design. Hydrogen sulphide is commonly present in O&G drilling and steel pipe, and
casings are available in grades that are resistant to hydrogen sulphide embrittlement.
From a corrosion standpoint, the casing material must be selected based on the project life span, the temperature
of the well, and the produced fluid chemistry. Depending on the conditions, the corrosion of the casing can be
managed by selecting materials that withstand elevated temperatures, such as high-chromium, nickel-based alloys
which can fortify casing integrity in geothermal environments. In lower-temperature but corrosive environments,
non-metallic liners can be run inside the casing; these provide excellent corrosion inhibition at much lower cost
compared to the advanced steel alloys. One example, NOV’s fiberglass TK Liner, has a maximum temperature of
only 121˚C; so, they can only be used in shallower well sections or in surface piping on an SHR project.
66
Figure 29 NOV’s Glass Reinforced Epoxy corrosion-resistant casing liner.
(NOV, n.d.).
Corrosion inhibition products are currently available at a TRL level 9, but in highly corrosive environments, those
products could strain the economics of SHR projects.
Conclusions
Corrosion is a common concern for several industries, and innovation into high-temperature corrosion inhibition
would benefit many industrial sectors, including geothermal. This innovation is best achieved through collaborative
research projects between universities, national laboratories, and commercial companies who can test materials
at full scale. Research into low-cost graphene coatings could be promising due to the conductive properties of
graphene and its high resistance to corrosion at high temperatures.
67
8. Conclusion
SHR geothermal systems have the potential to provide long-term, scalable, renewable baseload power. Unlocking
this potential requires significant innovation in drilling and well-construction technologies to improve ROP and
develop high-temperature electronic downhole tools for MWD, logging-while-drilling LWD, and directional drilling
in SHR conditions
Collaboration between the public and private research community can help create these R&D conditions by first
identifying all facilities around the world capable of SHR experimentation (e.g., Utah FORGE, IDDP), then creating
the necessary incentives for greater cooperation between major drilling companies and research groups, and
finally ramping up experimentation and R&D in SHR conditions.
Across all well-construction technology domains, one theme is clear: the technology to complete superhot or
ultradeep geothermal boreholes and wells exists—but we must reduce the overall cost and time on task to drill a
deep, superhot geothermal well. With continued support and development, the future of geothermal energy is
extraordinarily bright. A list of technology-specific findings and a gap analysis table are all included in the
Appendix.
68
Sources
Agrawal, M. G., Bernard, M. J., Hedegaard, M. M., Makthoum, M. M., Soni, M. A., & Wohrer, M. P. (2016). A
FRAMEWORK FOR INTERNATIONAL COLLABORATION ON LUNAR MISSIONS.
Ahmed, R. M., & Teodoriu, C. (2023). Applications of Radial Jet Drilling Techniques in Geothermal Wells.
Proceedings: 48th Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford, CA.
Ajima, K., & Naganawa, S. (2022). Evaluating wellbore cooling effect with insulated drill pipes in supercritical
geothermal well drilling. 46, 28–31.
Anders, E., Voigt, M., Lehmann, F., & Mezzetti, M. (2017). Electric impulse drilling: The future of drilling technology
begins now. 57762, V008T11A024.
Ando, R., & Naganawa, S. (2020). Simulating effect of insulated drillpipe on downhole temperature in supercritical
geothermal well drilling. 24, 26.
Blöcher, G., Peters, E., Reinsch, T., Petrauskas, S., Valickas, R., & van den Berg, S. (2016). D3. 2 Report on radial jet-
drilling (RJD) stimulation technology.
Bromley, C., Axelsson, G., Asanuma, H., Manzella, A., & Dobson, P. (2020). Supercritical Fluids–Learning about the
Deep Roots of Geothermal Systems from OEA Geothermal Collaboration. 1.
Bruno, M. S. (2005). Fundamental research on percussion drilling: Improved rock mechanics analysis, advanced
simulation technology, and full-scale laboratory investigations. Terralog Technologies Inc.
Champness, T., Worthen, T., & Finger, J. (2008). Development and Application of Insulated Drill Pipe for High
Temperature, High Pressure Drilling. Drill Cool Systems Incorporated.
Cladouhos, T. T., & Callahan, O. A. (2023). Heat Extraction from SuperHot Rock-Technology Development.
Denninger, K., Eustes, A., Visser, C., Baker, W., Bolton, D., Bell, J., Bell, S., Jacobs, A., Nagandran, U., & Tilley, M.
(2015). Optimizing geothermal drilling: Oil and gas technology transfer. Transactions, Geothermal
Resources Council.
Depouhon, A., Denoël, V., & Detournay, E. (2015). Numerical simulation of percussive drilling. International
Journal for Numerical and Analytical Methods in Geomechanics, 39(8), 889–912.
Drillstar - MUDHammer: High-power for hard rock drilling. (n.d.). Drillstar Industries. Retrieved March 11, 2024,
from https://www.drillstar-industries.com/drilling/mudhammer/
Dupriest, F., & Noynaert, S. (2022). Drilling practices and workflows for geothermal operations. IADC/SPE
International Drilling Conference and Exhibition.
El-Sadi, K., Gierke, B., Howard, E., & Gradl, C. (2024). Review Of Drilling Performance In A Horizontal EGS
Development.
Epplin, D. (2015). Drilling Motors for 300°C Geothermal Wells. 2015 IADC Extreme Drilling Technology Forum.
https://iadc.org/wp-content/uploads/2015/11/Dave-Epplin-DEC-Presentation.pdf
Ezzat, M., Adams, B. M., Saar, M. O., & Vogler, D. (2021). Numerical modeling of the effects of pore characteristics
on the electric breakdown of rock for plasma pulse geo drilling. Energies, 15(1), 250.
Fercho, S., Matson, G., McConville, E., Rhodes, G., Jordan, R., & Norbeck, J. (2024). Geology, Temperature,
Geophysics, Stress Orientations, and Natural Fracturing in the Milford Valley, UT Informed by the Drilling
Results of the First Horizontal Wells at the Cape Modern Geothermal Project.
Finger, J., Jacobson, R., Whitlow, G., & Champness, T. (2000). Insulated drill pipe for high-temperature drilling.
Sandia National Laboratories, Albuquerque, NM (US); Sandia National ….
69
Friðleifsson, G. Ó., Albertsson, A., Stefánsson, A., Þórólfsson, G., Mesfin, K. G., Sigurðsson, K., Sigurðsson, Ó., &
Gíslason, Ş. (2019). The Reykjanes DEEPEGS Demonstration Well–IDDP-2.
GA Drilling Plasmabit. (n.d.). GA Drilling. Retrieved March 11, 2024, from https://www.gadrilling.com/plasmabit/
Gerbaud, L., Pubill, A., Adannawy, H., Rouabhi, A., Lamazouade, M., Souchal, R., & Cuillier, B. (2022). Mud Hammer
Drilling in Deep Hard Geothermal formation, from the single impact to the directional and vibration
characterization. European Geothermal Congress 2022.
Graham, P., Krough, B., Nelson, T., White, A., & Self, J. (2017). Innovative Conical Diamond Element Bit in
Conjunction With Novel Drilling Practices Increases Performance in Hard-Rock Geothermal Applications,
California [C]. 223–227.
Holtzman, B., Groebner, N., Mittal, T., & Skarbek, R. (2023). Quench-Spallation Drilling: A Novel Drilling Head
Designfor Routine Heat Mining Above the Brittle-Ductile Transition. 6–8.
Houde, M., Araque, C., Oglesby, K., & Woskov, P. (2021). Rewriting the Limits for Deep Geothermal Drilling: Direct
Energy Drilling Using Millimeter Wave Technology. Proceedings World Geothermal Congress 2020+1,
Reykjavik, Iceland.
Houde, M., Woskov, P., Lee, J., Oglesby, K., Bigelow, T., Garrison, G., Uddenberg, M., & Araque, C. (2021).
Unlocking Deep SuperHot Rock Resources Through Millimeter Wave Drilling Technology. GRC
Transactions, 45.
Jahangir, E., Sellami, H., & Gerbaud, L. (2022). Improvement of drilling performances in deep geothermal drilling
in hard rocks, by using a novel Hydro-Mechanical Drilling. European Geothermal Congress 22.
Jerby, E., Aktushev, O., & Dikhtyar, V. (2005). Theoretical analysis of the microwave-drill near-field localized
heating effect. Journal of Applied Physics, 97(3).
Kaldal, G. S., Thorbjornsson, I., Gautason, B., Árnadóttir, S., Einarsson, G. M., & Reinsch, T. (2019a). Preparation
for Radial Jet Drilling Stimulation in a Geothermal Well in Iceland. European Geothermal Congress.
Kaldal, G. S., Thorbjornsson, I., Gautason, B., Árnadóttir, S., Einarsson, G. M., & Reinsch, T. (2019b). Preparation
for Radial Jet Drilling Stimulation in a Geothermal Well in Iceland. European Geothermal Congress.
Khaled, M. S., Wang, N., Ashok, P., & van Oort, E. (2023). Downhole heat management for drilling shallow and
ultra-deep high enthalpy geothermal wells. Geothermics, 107, 102604.
Khaled, M. S., Wang, N., Ashok, P., & van Oort, E. (2024). Downhole Temperature Estimation in Geothermal Wells
Using a Deep Learning Model Based on LSTM Neural Networks. D031S021R002.
Kukkonen, I., & Pentti, M. (2021). St1 Deep Heat Project: Geothermal energy to the district heating network in
Espoo. 703, 012035.
Kukkonen, I. T., Heikkinen, P. J., Malin, P. E., Renner, J., Dresen, G., Karjalainen, A., Rytkönen, J., & Solantie, J.
(2023a). Hydraulic conductivity of the crystalline crust: Insights from hydraulic stimulation and induced
seismicity of an enhanced geothermal system pilot reservoir at 6 km depth, Espoo, southern Finland.
Geothermics, 112, 102743.
Kukkonen, I. T., Heikkinen, P. J., Malin, P. E., Renner, J., Dresen, G., Karjalainen, A., Rytkönen, J., & Solantie, J.
(2023b). Hydraulic conductivity of the crystalline crust: Insights from hydraulic stimulation and induced
seismicity of an enhanced geothermal system pilot reservoir at 6 km depth, Espoo, southern Finland.
Geothermics, 112, 102743.
Lehmann, F., & Reich, M. (2015). Development of Alternative Drive Concepts for Down-the-Hole Hammer in Deep
Drilling. 56581, V010T11A007.
70
Liu, W., Zhang, Y., Zhu, X., & Luo, Y. (2023). Influence of Pore Characteristics Inside Formation on Partial Electrical
Breakdown in Electric Impulse Drilling. IEEE Transactions on Plasma Science.
Mohamed, A., Salehi, S., & Ahmed, R. (2021). Significance and complications of drilling fluid rheology in
geothermal drilling: A review. Geothermics, 93, 102066.
Moore, J., McLennan, J., Allis, R., Pankow, K., Simmons, S., Podgorney, R., Wannamaker, P., Bartley, J., Jones, C.,
& Rickard, W. (2019). The Utah Frontier Observatory for Research in Geothermal Energy (FORGE): An
international laboratory for enhanced geothermal system technology development. Proceedings.
Moore, J., McLennan, J., Pankow, K., Finnila, A., Dyer, B., Karvounis, D., Bethmann, F., Podgorney, R., Rutledge, J.,
& Meir, P. (2023). Current activities at the Utah Frontier Observatory for Research in Geothermal Energy
(FORGE): A laboratory for characterizing, creating and sustaining Enhanced Geothermal Systems. ARMA-
2023-0749.
Moore, J., McLennan, J., Pankow, K., Simmons, S., Podgorney, R., Wannamaker, P., Jones, C., Rickard, W., & Xing,
P. (2020). The utah frontier observatory for research in geothermal energy (Forge): A laboratory for
characterizing, creating and sustaining enhanced geothermal systems. Proceedings of the 45th Workshop
on Geothermal Reservoir Engineering.
Naganawa, S., Tsuchiya, N., Okabe, T., Kajiwara, T., Shimada, K., & Yanagisawa, N. (2017). Innovative drilling
technology for supercritical geothermal resources development. Proceedings.
Nair, R., Peters, E., Šliaupa, S., Valickas, R., & Petrauskas, S. (2017). A case study of radial jetting technology for
enhancing geothermal energy systems at Klaipeda geothermal demonstration plant. 13–15.
Oglesby, K., Woskov, P., Einstein, H., & Livesay, B. (2014). Deep geothermal drilling using millimeter wave
technology (final technical research report). Impact Technologies LLC, Tulsa, OK (United States).
Orchyd Project. (n.d.). Orchyd. Retrieved March 11, 2024, from https://www.orchyd.eu/project/
Petty, S., Uddenberg, M., Garrison, G. H., Watz, J., & Hill, B. (2020). Path to Superhot Geothermal Energy
Development.
Phuoc, T. X., Massoudi, M., Wang, P., & McKoy, M. L. (2020). A study of temperature distribution and thermal
stresses in a hot rock due to rapid cooling. Journal of Heat Transfer, 142(4), 042302.
Pink, A., Patterson, A., & Thoresen, K. E. (2023). Building a System to Solve the Challenges of Drilling Hot Hard Rock
for Geothermal and Oil and Gas. SPE/IADC International Drilling Conference and Exhibition.
Pink, T., Shiller, J., Pink, S., & Carlson, A. (2022). The success of world’s first Hybrid PDC and Particle Drilling bit for
Geothermal Applications.
Pollack, A., Horne, R., & Mukerji, T. (2020). What are the challenges in developing enhanced geothermal systems
(EGS)? Observations from 64 EGS sites. 1.
Rahmani, R., Pastusek, P., Yun, G., & Roberts, T. (2021). Investigation of PDC cutter structural integrity in hard
rocks. SPE Drilling & Completion, 36(01), 11–28.
Reinicke, A., Deb, P., Saar, M. O., Zikovic, V., Lassnig, E., Knebel, M., & Jette Blangé, J. (2023). Novel Directional
Steel Shot Drilling Technology for Short-Radius Multilaterals-Field Application and Commercial Impact.
EGU-14928.
Reinsch, T., Dobson, P., Asanuma, H., Huenges, E., Poletto, F., & Sanjuan, B. (2017). Utilizing supercritical
geothermal systems: A review of past ventures and ongoing research activities. Geothermal Energy, 5(1),
1–25.
Roberts, S. L., Bailey, M. J., Victor, P., Sutan, M., Prasetia, B., & Lock, E. (2022). Building up steam, a progressive
approach to drilling volcanic geothermal formations. D021S035R006.
71
Self, J., Stevenson, M., Roberts, T., Rivera, R., & Onderko, L. (2021). Fixed cutter bit & cutter technology set new
performance standards for geothermal drilling. GRC Transactions, 45, 826–846.
Shryock, S. (1984). Geothermal well cementing technology. SPE-12454.
Song, H., Shi, H., Yuan, G., Xia, Y., Li, J., & Liu, T. (2022). Experimental study of the energy transfer efficiency and
rock fragmentation characteristics in percussive drilling. Geothermics, 105, 102497.
Stefánsson, A., Duerholt, R., Schroder, J., Macpherson, J., Hohl, C., Kruspe, T., & Eriksen, T.-J. (2018). A 300 degree
celsius directional drilling system. IADC/SPE Drilling Conference and Exhibition.
Stoxreiter, T., Rehatschek, K., & Hofstätter, H. (2019a). ThermoDrill—Development of an alternative drilling
technology for deep geothermal applications. World of Mining—Surface and Underground, 71(5), 276–
282.
Stoxreiter, T., Rehatschek, K., & Hofstätter, H. (2019b). ThermoDrill—Development of an alternative drilling
technology for deep geothermal applications. World of Mining—Surface and Underground, 71(5), 276–
282.
Sugiura, J., Lopez, R., Borjas, F., Jones, S., McLennan, J., Winkler, D., Stevenson, M., & Self, J. (2021). Oil and Gas
Drilling Optimization Technologies Applied Successfully to Unconventional Geothermal Well Drilling. SPE
Annual Technical Conference and Exhibition.
Tester, J. W., Anderson, B. J., Batchelor, A. S., Blackwell, D. D., DiPippo, R., Drake, E. M., Garnish, J., Livesay, B.,
Moore, M., & Nichols, K. (2006). The future of geothermal energy. Massachusetts Institute of Technology,
358, 1–3.
Thorsteinsson, H., Augustine, C., Anderson, B., Moore, M., & Tester, J. (2008). The impacts of drilling and reservoir
technology advances on EGS exploitation. 1–14.
van Oort, E., Chen, D., Ashok, P., & Fallah, A. (2021). Constructing deep closed-loop geothermal wells for globally
scalable energy production by leveraging oil and gas ERD and HPHT well construction expertise.
D021S002R001.
Visser, P. W., Henk, K., Bense, V., & Emiel, B. (2020). Impacts of progressive urban expansion on subsurface
temperatures in the city of Amsterdam (The Netherlands). Hydrogeology Journal, 28(5), 1755–1772.
Walsh, S. D., & Vogler, D. (2020). Simulating electropulse fracture of granitic rock. International Journal of Rock
Mechanics and Mining Sciences, 128, 104238.
Wang, H., Liao, H., Wei, J., Liu, Y., & Chen, J. (2021). Downhole Stress Release Effect of Grooving by Means of Ultra-
high Pressure Water Jet in Deep Geothermal Drilling.
Wang, H.-J., Liao, H.-L., Wei, J., Liu, J.-S., Niu, W.-L., Liu, Y.-W., Guan, Z.-C., Sellami, H., & Latham, J.-P. (2023). Stress
release mechanism of deep bottom hole rock by ultra-high-pressure water jet slotting. Petroleum Science,
20(3), 1828–1842.
Woskov, P. (2017). Millimeter-Wave Directed Energy Deep Boreholes. AIG DRILLING FOR GEOLOGY II CONFERENCE
Brisbane, Australia.
Woskov, P., & Cohn, D. (2009). Annual report 2009: Millimeter wave deep drilling for geothermal energy, natural
gas and oil MITEI seed fund program.
Woskov, P. P., Einstein, H. H., & Oglesby, K. D. (2014). Penetrating rock with intense millimeter-waves. 1–2.
Xiao, S., Ren, Q., Ge, Z., Chen, B., & Wang, F. (2020a). Study of the rock-breaking and drilling performance of a self-
rotatory water-jet bit in water-jet drilling and its influential factors. Energy Sources, Part A: Recovery,
Utilization, and Environmental Effects, 1–17.
72
Xiao, S., Ren, Q., Ge, Z., Chen, B., & Wang, F. (2020b). Study of the rock-breaking and drilling performance of a
self-rotatory water-jet bit in water-jet drilling and its influential factors. Energy Sources, Part A: Recovery,
Utilization, and Environmental Effects, 1–17.
Zhang, A., Millmore, S., & Nikiforakis, N. (2023). Thermal simulation of millimetre wave ablation of geological
materials. Computers and Geotechnics, 161, 105571.
Zhang, Q., Wang, G., Pan, X., Li, Y., He, J., Qi, Y., & Yang, J. (2023). High Voltage Electric Pulse Drilling: A Study of
Variables through Simulation and Experimental Tests. Energies, 16(3), 1174.
73
Appendix
Technology specific findings:
Conventional drilling
• Current status:
o The highest ROP and longevity performance with PDC bits in SHR conditions
o Advanced high-temperature, hard rock specific drill bits are commercially available.
o The most mature drilling technique with high availability of drill rigs and operators relative to
hybrid conventional methods
o High level of directional control throughout borehole
• Technology gaps:
o Design and optimization of cutters and bits to reach ultra-deep depths
o Ability to reach SHR conditions of 375˚C
o Weight on bit optimization for increased depths
Percussive drilling
• Current status:
o Percussive systems have been deployed in supercritical conditions
o 20 m/h ROP in hard rock formations at 6000 m
• Technology gaps:
o Air hammer drilling inefficiency at depths due to loss of pressure
o Need for real-work supercritical testing
• Current status:
o Under testing and simulation for super- and ultra-deep wells
o Reduces strain in rock formation therefore reducing tooling wear
• Technology gaps:
o Surface systems to be designed to maintain flow rates and required pressures
o Increased need for directional control systems
o Improved understanding of standoff distance from borehole bottom
o Still will require tripping and tooling repair for conventional bits
74
Plasma drilling
• Current status:
o Combined system of plasma and conventional is under evaluation; testing in approximately 18
months
o ROP in hard rock formations of up to 7.2 m/h in SHR conditions
o Designed to be plug-and-play with current conventional drilling surface systems
• Technology gaps:
o Efficiency of material removal is based on material properties, specifically boundary layers of the
formation Circulation system for required dielectric fluid under evaluation
o Maintaining power delivery to the drill bit
• Current status:
o Laboratory testing completed in hard rock simulations, maximum drilled distance of 8 m
o Short borehole field testing to be completed fiscal year 2024 - 25
o Theoretical benefits include avoidance of trip time
• Technology gaps:
o Creation of drilling rig compatible to a 1 MW gyrotron
o Ruggedization of the gyrotron
o Impact of the property’s hard rock formations and the effect on drilling, porosity, and dielectric
conditions
o Structural integrity of the waveguide at super- and ultra-deep depths
o Lack of field testing to validate performance
Particle drilling
• Current status:
o Laboratory and semi-scale testing completed with claimed market readiness in fiscal year 2024
o Downhole hardware not affected by temperature and pressure for SHR drilling
o Applicable for long runs in hard rock formations
• Technology gaps:
o Inability to support direction control due to impacts with injected particles
o Maintaining nozzle performance in downhole operations to continue circulation of injected
particles
• Current status:
o High-temperature tools have been manufactured, tested, and run regularly at 175 ºC
75
o A smaller range of tools manufactured by Baker Hughes, Hephae, and MB Century are rated to
withstand >200 ºC
• Technology gaps:
o Rare materials are required to build electronics that can withstand SHR temperatures but they
are expensive and slow to manufacture
o An ideal system incorporating existing high-temperature tools (including one-run, encapsulated
or standalone) with temperature management equipment must be modelled and tested in the
field
• Current status:
o Components that support 400˚C currently available: insulated drill piping, coatings, and mud
chillers
o Industry developing new and novel techniques to support SHR applications
• Technology gaps:
o Current carbon supply insufficient for geothermal at scale
o Integration and packaging of systems still required
o Possible introduction of exotic materials to improve efficiency
Corrosion inhibition
• Current status:
o Corrosion inhibiting products are available but expensive and possible providing a negative impact
to geothermal projects
o Understanding of corrosion present in hydrocarbon industry, providing expert knowledge to
improve systems for geothermal applications
• Technology gaps:
o Managing location specific material to accommodate location-based ground composition
76
Technology gap summary table
Technology Criticality
Future required state of
Technology Current state of technology readiness for FOAK Steps needed to get to TRL 9
technology
level (TRL) project
Conventional Roller cone bits rated to 288oC We need to see ever PDC bits are at 1 Conventional PDC bits are
drill bits, PDC, performed successfully at high improving performance, TRL 8-9, they already at a TRL 9 for SHR but
roller cone, and temperatures on the IDDP increased durability, faster just need to be because the performance of
diamond high temperature projects in rates of penetration, and tested and run these products has such a
impregnated Iceland but they should be higher temperature at the SHR large impact on project
phased out because of low ratings. New bit designs temp level. economics innovation must
ROPs. PDC drill bits have and cutters are required Roller cones continue at pace. The biggest
shown massive performance for interbedded formations are at TRL 8-9 impact on performance
jumps on the FORGE project with large differences but have comes from rapid iteration
and most recently on the hardness between the probably which comes from drilling
FERVO location. Current different layers. Fast ROPs reached their lots of wells. We need to drill
designs are delivering runs of in basalts and very high performance 10 x more geothermal wells
more than 900 M at 30 m/hr compressive strength limits. in a lot of different rock
in hard crystalline granite. rocks. Increased stability to lithologies then speed and
Performance in igneous rocks be able to handle vibration durability will increase
with layers of high differential in the hard basement rapidly. All the drill bit
strength is still relatively rocks. We will need to see manufacturers are working
unknown but performance in a proven track record of on material science of the
New Zealand with large PDC PDC performance in bodies and advanced
bits was very promising. The igneous and metamorphic diamond technology for the
current state of PDC rocks above 400 oC. We cutters. Some collaboration
performance has already need to see the impact of between universities, labs
significantly changed the thermal spallation in and manufacturers would
economics of SHR projects by combination with PDC bits, probably speed up future
reducing the number of days the lab results look development. Studies are
on location significantly. Drill promising. needed, on hybrid
bits are all steerable and the conventional bits coupled
current temperature with other secondary
limitation of PDC bits is the techniques like thermal
“O” Rings holding in the spallation, particle impact,
nozzles which is close to plasma, and water jet.
400oC.
Percussive Percussive drilling combines Percussive drilling is a Percussion 2 Further experimentation at
air or mud hammer drilling mature drilling method drilling for SHR SHR test sites will validate the
with rotary drilling techniques, that requires repeated is at a TRL 7 performance of percussion
that have demonstrated the deployments in SHR status. They drilling for SHR projects. Like
capacity to drill to SHR conditions to verify its high have rotary methods, diversity in
conditions, or deep into hard ROP can be sustained in demonstrated geological environments, drill
rock. The leading firms behind this domain. Their high ROP and bits and manufacturers will
percussive drilling include performance must match bit longevity in identify the optimal approach
Strada Global, Orchyd, and or surpass conventional hard rock but to reach SHR with percussive
HydroVolve. Percussive drilling rotary methods if it is to require further systems. This could be done
was deployed for Stn-1, Fervo become the primary testing in SHR coevally with rotary system
and IDDP-2. High ROP rates, method of SHR drilling, but domains. experimentation at SHR sits,
20 m/hr, have been this can only be achieved so both systems are directly
demonstrated to >6000 m, with aggressive repetitions comparable. Coordination
suggesting that percussive of in-field deployments. between percussive drilling
drilling could be competitive firms may improve
with conventional rotary innovation rates, but this
methods in ROP generation requires strategic sharing of
through hard rock. Further proprietary information.
testing in SHR conditions is
required.
Water jet Water jet drilling is under two An increase in Water jetting 2 Water-based systems are not
different application types: 1) understanding of the is TRL 5 for yet TRL 9. This is due to
Radial jetting, which has been performance, specifically in deep multiple factors in regard to
successfully applied to highly compressive geothermal designs of the systems and
stimulate current geothermal formations, will be critical drilling in hard understandings of interaction
systems and 2) the use of to deep borehole water- rock. in hard rock for SHR
78
high-pressure systems, along jetting applications as well applications. These items
with a conventional drill it, to as ensuring systems are include, better directional
remove material downhole able to maintain the high control of fluid systems,
and reduce the stresses fluid flow rates required. increased modeling of flow
observed by the conventional For radial systems, it will be dynamics on hard rock
drilling system. The Orchyd more in line with system formations, and performance
Project is the leader in water- optimizations as it has in high compressive strength
based systems with their proven in the stimulation formations.
design to combine water- of under-performing
jetting and percussive in a geothermal systems.
singular system.
Plasma Plasma drilling is currently in Real-world testing needs to Plasma based 4 To Reach TRL 9, systems will
the design phase. GA drilling, be completed to ensure systems are need to be tested in SHR
in conjunction with Nabors technology application in currently at a conditions to validate
Industries, are developing SHR conditions. New TRL 3-4 since performance. Currently,
both a plasma-conventional surface level material and deep borehole there is a lack of
combined system, with fluid handling systems validation has understanding of what the
downhole power generation need to be designed to not been true performance of the
as well as a stand-alone work in conjunction with completed. systems will be. Design of
plasma system, most probably current hydrocarbon electrical systems that do not
run on coiled tubing. The systems. Increased pose any interference to
current timeline is to have the understanding of required systems will also
combined system completed downhole power need to be validated.
and begin testing within 18 requirements and the
months. effects of geological
boundary layers will be
critical to increasing
effectiveness of the overall
system.
MMW MMW Drilling is in its design Scale up in the lab and in- MMW drilling 4 To reach TRL 9, repeated in-
and laboratory field is required to verify is at a current field experiments
experimentation phase, the potential to reach SHR status of TRL demonstrating scale up
79
spearheaded by Quaise, domains with MMW. This 2-3. This may potential for SHR drilling is
Nabors, ARPA-E and ONRL. initiates with acquiring a be advanced required. Areas of inquiry
They have proof-of-concept high powered gyrotron, when in field include validation of the
that MMW drilling can with >1MW capacity. tests are wave guide deployment, rig
vaporize rock to create a well, Equipment ruggedization successfully compatibility of the gyrotron,
however the maximum depth and system design, as well completed. and the efficiency of rock
penetration thus far is 2.4 m as operability for drilling vaporization. The design of a
by 2 cm diameter. Plans to personnel follows this step. drilling stand capable of
scale up experiments in the Further modelling of mmw simulating SHR conditions is
lab and in-field are in place, as absorptive properties for necessary for drilling
is the acquisition of a high heterogeneous rocks is also experimentation.
powered, 1 MW, gyrotron that required.
may advance the penetration
depth. Transmission of the
mmw downhole is achieved
with metal waveguide that
must maintain a near-perfect
vertical orientation or be at
risk of vaporization. Methods
to deploy the waveguide and
cope with this constraint are
under development. Bottom
hole conditions are monitored
through the waveguide,
through reflectometry,
spectroscopy, and radar,
although this concept requires
further testing.
Particle impact Particle impact drilling is The latest design Particle 1 Particle Drilling is
drilling provided by 2 companies, modifications, post FORGE, Drilling are at (Particle commercial, but the
Particle Drilling (PD) and should deliver ROPs over TRL 5 and Drilling) company needs a buyer or a
Canopus. PD injects 3% 30 M/hr and a run length 2 significant investor to be able
80
particles (2 mm diameter) into of 600M. For particle Canopus are (Canopus to continue operating. It then
the flow line and then drilling to compete with TRL 4. ) needs to be deployed to a
accelerates those particles to conventional PDC bits the SHR test rig and or a
impact and destroy the brittle system needs to deliver run commercial project and
crystalline rocks. The PD drill lengths over 1000 m. iterate over multiple wells.
bit is a hybrid PDC reamer Particle Drilling has a lot of Compatibility with MWD
with particle accelerators. The potential in very hard tools needs to be assessed
Canopus system is somewhat interbedded rocks with lots and if required integration
similar, but it fires smaller (1 of layers of different with a negative pulse tool or
mm Diameter) particles out hardness. Canopus needs wired drill pipe.
through the cutting face of the to full scale test and Canopus needs a full-scale
PDC bit. The Canopus system manufacture a set of tools test on a vertical drilling rig
is also capable of steering the to run on a pilot project. and can ruggedize and
well and comes with and They also need to be optimize the whole system.
integrated MWD and RSS tool. competitive with Their current MWD is an EM
conventional PDC bits from tool which has significant
ROP and durability. limitations in deep hard
crystalline rock. They need to
investigate an alternative
telemetry method like
negative pulse or wired pipe.
Both Canopus and Particle
Drilling would benefit from a
significant government grant
and or a large private
investor. Both technologies
show huge promise and are
very close to being
commercial.
High - A wide range of high It is possible to build 400 ºC TRL 8 when 2 To get to TRL 9 a complete
temperature temperature tools have been electronics out of exotic used with system needs to be modeled,
tools manufactured, tested, and run materials like Silicon manufactured, tested, and
81
regularly at 175 ºC. There is a Carbide and Gallium nitride other cooling run in the field to be field
smaller range of tools at 200 but the practical methods. hardened. The system will
ºC. One manufacturer application of this over the TRL 7 if then need to be iterated to
(Hephae) is working on 200- next 20 years is encapsulated. reduce cost and improve
300 ºC with the first iteration challenging. The preferred TRL 3 for 400 performance.
at 210 ºC. Baker Hughes has solution will be to build the ºC stand-alone
run insulated tools with a most economically tools.
cooling system above 300 ºC practical solution
on IDDP-2. MB Century have combining other
wireline tools for logging the temperature management
hole up to 350 ºC methods and the highest
rated electronic tools.
Insulated drill There are three insulated drill The required state is that The (Drill Cool) 2 To get to TRL 8 a complete
pipe pipe products and concepts. the insulated pipe lowers system has system needs to be modeled,
External coated (Drill Cool) has the fluid BHCT while drilling been run but manufactured, tested, and
been manufactured and run to below 175ºC and return not at SHR run in the field. A partnership
on 2 projects. Internal coated temperature to below conditions, so between a SHR project owner
(NOV) has been modelled and 100ºC. Modelling has TRL 7 is most and a pipe manufacturer
designed, test joints are being showed that this can be appropriate. needs to be formed. This
built. Vacuum sealed achieved with a partial process could be
(Vallourec) has been built and string of insulated pipe considerably accelerated with
run for production purposes used in combination with a government grant or
but currently not suitable for other temperature private investment.
drilling management technologies.
Drilling fluid High temperature drilling Multiple fluid systems A SHR water- 2 Labs and Universities
fluids with reasonable should be designed that based system combined with commercial
rheology have been run up to are fit for purpose for SHR is at a TRL 8 partnerships should be
353 ºC. Water can be used at conditions. These fluids for low angle working on water-based
SHR temperatures but in its need to be able to cool the holes and a additives to create a water-
simplest form it may only be borehole, lubricate the TRL 7 for based system that has all the
fit for purpose to clean the hole, transport cuttings higher angle rheological characteristics to
hole at low inclinations. back to the surface and in or horizontal be able to efficiently drill and
82
Supercritical, liquid, and some cases minimize fluid wells. Using clean a SHR well.
gaseous CO2 has potential as a losses or reduce the risk of CO2 as a fluid Government grants, private
working fluid, but it is in the fluid losses in the well. system is at equity investment could
concept, modelling design Separate systems will need TRL 3-4. accelerate the development
phase. to be designed where of CO2 as a working fluid and
consideration is made for reduce the time to market by
impermeable rock with low several years.
fluid loss risk and a
permeable or fractured
rock with high fluid loss
risk. High temperature 400
ºC test equipment is
required.
Low heat Low heat coefficient coatings, Low heat coefficient The current 0 Development of even higher
coefficient applied to conventional drill coatings are a very coating is TRL capability coatings could be
coatings pipes, as part of a wellbore practical, low-cost way of 9 but needs to accelerated by partnerships
cooling strategy are reducing the temperature be combined between commercial entities,
commercially available today of the fluid going into the with other labs, and universities. There
(NOV) and have been run into wellbore. The lower the temperature are probably materials out
high temperature wells. They thermal conductivity of this management there that could make a
have not been run or tested at coating the more effective equipment to significant jump in the
SHR temperatures. it is at maintaining a lower deliver the performance of these
temperature fluid. Work is required fluid coatings. R&D into materials
required to produce a thin, temperature science could yield a
effective coating with even at the tools breakthrough technology as a
higher capability than 0.16 and the bit. coating or part of a coating.
W/mk.
Mud coolers Mud coolers for SHR projects A highly effective mud Mud coolers 0 Work can be done by the
are commercially available cooling system that can are commercial entities to
today. There are multiple manage a high volume of commercially develop coolers that can
manufacturers and vendors. high temperature fluid and available handle a higher volume of
Their effectiveness is reduce the temperature of today and are fluid and deliver a large in
83
somewhat dictated by the rest that working fluid TRL 9. They and out temperature
of the temperature significantly so it can be can be used in differential at increased flow
management system and returned down the string parallel or in rates. Research into complete
environmental conditions. at the appropriate series to temperature management
temperatures. That improve and the sensitivities of
temperature needs to be effectiveness. various components should
low enough so the other They do have be done to try and find the
components in the a relatively most cost-effective solution.
temperature management high footprint
system can keep the on the
temperature at the bit location. They
below 175 ºC may require a
large volume
of fresh water
to operate
which may
cause
challenges.
Super critical CO2 as a working fluid for A full prototype CO2 system Using CO2 as a 3 A SHR test location needs to
CO2 geothermal wells has been needs to be built and working fluid be developed somewhere in
investigated since around tested on test location that for SHR the US and potentially
2000. This work has been has SHR conditions. conditions is elsewhere in Europe, Africa,
mainly theoretical R&D. The TRL 3-4 or Asia. The quickest way to
physics looks very promising do this would be potentially
and there are oil and gas drill a deeper low angle well
service companies who have at the FORGE location or
the capability of building a create a similar location in
complete system. One of the the western USA. A
challenges could be sourcing partnership needs to be
plentiful volumes of CO2 at low formed between a
cost. commercial entity who can
develop this technology and
84
universities and national labs
to support. Significant DOE or
EU grants could accelerate
development but access to a
SHR test well is essential.
Fluid Dynamics, It is well-understood that the A real-time SHR hydraulics TRL 7. 1 This work could be done by a
flow rates, creation of laminar flow in the model needs to be created Hydraulics university, national lab and or
turbulent vs. annulus of the borehole can where the borehole models exist a commercial entity. A small
Laminar flow have a significant positive temperature can be but they need philanthropic grant or a DOE
effect on insulating the fluid in modelled and adjusted in to be adapted grant could accelerate the
the drill string. real-time to be able to to SHR process.
create an environment conditions and
where the temperature of be turned into
the fluid is optimized. a real-time
application.
Coatings, Corrosion inhibiting products Low-cost corrosion Casing and 2 Learning from other
Material for SHR casing are available solutions using coatings is production industries like the
Science, today, but they are very the ideal solution for the ID tubing is Petrochemical industry could
Advanced expensive. steel alloys using of the drill pipe. Drill pipe available have a significant benefit
alloys, Non chromium and other metals alloys can also be selected today that can here. Otherwise,
Metallics, are available that can survive to minimize the risk of manage collaboration between labs,
Graphene, in almost any environment. Hydrogen embrittlement corrosive, universities and industrial
Ceramics Corrosion needs to be taken when H2S is encountered. high- partners is needed to find
into consideration by the The future state would be temperature some lower cost solutions to
drillers during the casing finding commercially viable environments, the corrosion problems in
design process, but this topic corrosion solutions for all but it is SHR drilling.
will be covered in more detail steel tubulars that go into expensive. TRL
in other papers. Corrosion of the SHR environment. 9. Drill pipe
the drill pipe does need to be alloys and
considered and drill pipes strengths are
need to be specified to handle available that
these environmental can handle
85
conditions. The easiest way to corrosion in
manage corrosion is to drill SHR
SHR CLG and EGS wells where conditions.
the drilling fluid chemistry can Best solution
be managed on the surface. is to drill hot
dry rock and
avoid
hydrothermal
formation
fluids that are
highly acidic.
A very limited supply of rigs The main problem here is TRL 9 1 A test location or commercial
that are available today that not the technical readiness equipment is SHR operation is needed that
have all the specifications that but the limited supply. available has a long-term schedule of
are required to drill SHR wells. Most of the rigs that are today even if it wells (Greater than 5 years).
The longest hydrocarbon well suited for these projects is not quite A project like this would
to date is 15240 and was are either old and require optimal. provide an incentive for a rig
drilled off the Orlan Platform significant overhauls or are manufacturer to build a
in Abu Dhabi. There are a few on long-term Oil and Gas custom fit for purpose rig for
land rigs like the Doyon 26 and projects. If new rigs are SHR geothermal. DOE funding
the Deutag T45 that have the being built for SHR of such a project would
pump capacity, the fluid geothermal projects then accelerate the development
storage capability, top drive additional consideration of these rigs and the supply
torque, derrick strength, sub needs to be made for chain that needs to be
structure, BOPs and drilling in more densely created for such a facility.
drawworks capability to drill populated areas, taking
these wells. The downhole power from overhead lines,
temperature will need to be noise reduction, robotics,
managed to ensure fluid small loads, smaller crews.
return temperature is below
boiling point and does not
damage BOP rubber elements.
86
The rigs may need to come
with additional equipment like
MPD, mud coolers, particle
drilling, automation, robotics,
and continuous circulation.
Additional pad space may be
required to handle large
volumes of fluid to allow
evaporative cooling.
87