Вы находитесь на странице: 1из 6

DISTINGUISHED AUTHOR SERIES

Hydrates: State of the Art Inside and Outside Flowlines


E.D. Sloan, C.A. Koh, A.K. Sum, A.L. Ballard, G.J. Shoup, N. McMullen, J.L. Creek, and T. Palermo

Abstract The state of the art of three hydrate applications in petroleum engineering is presented in order of decreasing importance: (1) flow assurance, (2) energy resource, and (3) climate change. In flow assurance, there is a hydrate-plug-prevention shift under way: from avoidance to management of hydrate formation. In addition to avoiding the region of hydrate stability by injecting thermodynamic inhibitors, time-dependent studies enable flow-assurance engineers to better address such concerns as flowline restarts, cold (stabilized) flow, low-dosage hydrate inhibitors, and plug remediation. These applications are related to conceptual ideas of hydrate-plug formation in oil and condensate systems. The second area, energy resources, is marked by a transition to an extended production test in the permafrost, and to characterizing resources and economics in the marine environment. The third area, climate change caused by hydrates, is an area of current research. Preliminary estimates suggest no abrupt methane contribution to the environment from hydrates in the immediate future. Introduction Natural-gas hydrates (clathrates) are crystalline, ice-like solids that form when small (<9 ) molecules contact water at low temperature and high pressure (Sloan and Koh 2008). Because hydrates are solids, they can block oil and gas flow-

lines; flow assurance involves considerable effort and expense to ensure that this blockage does not happen. Also, hydrates concentrate hydrocarbon gases by a volumetric factor of 164 (at standard temperature/pressure), such that large deposits of natural hydrates are becoming factors in energy resources and climate concerns in marine and permafrost deposits. A combination of applications, principally in flow assurance, energy, and, more recently, climate change, has driven hydrate engineering and science for 75 years. The number of published hydrate-related papers is an exponential function of time, extrapolating to more than two refereed publications per day in this decade. As Table 1 summarizes, triennial international hydrate conferences have been held since 1993; the most recent conference, in July 2008, presented more than 400 manuscripts. From such conferences, it is possible to determine the state-of-the-art in hydrates, both inside and outside the flowline. This work deals principally with flow assurance as the main hydrate concern of petroleum engineers, followed by a brief introduction to hydrates as an energy resource, before concluding with a few statements on climate change. Inside the Flowline: Flow Assurance In a survey of 110 energy companies, flow assurance was listed as the major technical problem in offshore energy development. Hydrate plugs are the largest concern by an order of magnitude because hydrate plugs form so rapidly and without warning in offshore lines relative to waxes, scales, or asphaltenes. In contrast to the hydrate-plug-formation rates, remediation can require days or months. For flow assurance, one major task is to develop a plan to manage potential problems of hydrate formation, along with dealing with plugs when they occur. This task includes the ability to predict where and approximately when hydrate plugs might form, and to prevent hydrate-plug formation. When all of the water cavities in hydrates are occupied, all common hydrate structures contain approximately 15 mol% hydrocarbon as guests, and 85% water as hostsunusually high concentrations for compounds so dissimilar. Because the solubility of methane in liquid water (1 methane molecule in
Copyright 2009 Society of Petroleum Engineers This is paper SPE 118534. Distinguished Author Series articles are general, descriptive representations that summarize the state of the art in an area of technology by describing recent developments for readers who are not specialists in the topics discussed. Written by individuals recognized as experts in the area, these articles provide key references to more definitive work and present specific details only to illustrate the technology. Purpose: to inform the general readership of recent advances in various areas of petroleum engineering.

Dendy Sloan, SPE, holds the Weaver Chair in Chemical Engineering and is the Director of the Center for Hydrate Research at the Colorado School of Mines (CSM). He was an SPE Distinguished Lecturer 199697. Carolyn Koh is Associate Professor and Codirector of the CSM Hydrate Center. Amadeu K. Sum is an Assistant Professor in the Chemical Engineering Department at CSM and Codirector for the CSM Center for Hydrate Research. Adam L. Ballard is Facility Engineering Team Leader for BPs Thunder Horse production platform in the deepwater Gulf of Mexico. George Shoup is the Paleogene Facilities Subsea Engineering Manager for BP . Norm McMullen, SPE, is a retired Senior Advisor from BP plc and is a Flow Assurance Consultant. Jeff Creek, Chevron Oil Field Research Company, works with the Flow Assurance Core Team in fluid analysis and phase behavior as integrated with multiphase flow. Thierry Palermo works in the production department at Total E&P .

JPT DECEMBER 2009

89

DISTINGUISHED AUTHOR SERIES

TABLE 1HISTORY OF TRIENNIAL INTERNATIONAL CONFERENCES ON GAS HYDRATES Number of Papers/ Authors 61/130 87/195

Hydrateformation curve

f
Hydratefree region

Date June 1993 June 1996

City/Sponsor(s) New York/New York Academy of Sciences Toulouse/cole Nationale Suprieure d'Ingnieurs de Gnie Chimique Salt Lake City/New York Academy of Sciences Yokohama/Keio University Trondheim/Statoil

July 1999 May 2002 June 2005 July 2008

104/258 204/500 247/330

Fig. 1Hydrate-formation pressures and temperatures as a function of methanol concentration in free water for a given gas mixture. Flowline-fluid conditions are shown at distances along the bold black curve (Notz 1994).

Vancouver/University of 417/500+ British Columbia and National Research Council, Canada

1,200 water molecules) and the solubility of water dissolved in methane gas (1 water molecule in 1,000 methane molecules) are low compared to their high concentrations in hydrate, hydrate forms most easily at phase interfaces where there is an abundance of both hydrocarbon and water. Specifically, for gas+water systems, hydrate forms at the gas/water interface. However, for gas+water+oil systems, hydrate typically forms at the interface between the two liquids, water and oil, from small gas molecules dissolved in the oil. This interfacial phenomenon is a key concept in understanding hydrate formation and prevention. Example: Hydrates in Flow Assurance. A case study of hydrate formation is shown in the pressure/temperature diagram of Fig. 1 for a deepwater flowline fluid. To the right, at high temperature and moderate pressure in the diagram, hydrates will not form and the system will exist in the fluid (hydrocarbon and water) region. However, hydrates will form in the shaded region at lower temperatures marked Hydrate-Forming Region, so hydrate-prevention measures should be taken. In Fig. 1, at 7 miles from the subsea wellhead, the flowing stream retains some reservoir heat, thus preventing hydrates. The ocean cools the flowing stream, and at approximately 9 miles, a unit mass of flowing gas and associated water enters the hydrate region to the left of the hydrate-formation curve, remaining in the uninhibited hydrate area until mile 45. By mile 30 the temperature of the pipeline system is within a few degrees of the deep ocean temperature (40F). To prevent hydrate formation and blockage, approximately 23 wt% methanol, with some safety factor, is required in the free-water phase to shift the hydrate-formation region to the left of flowline conditions, to prevent hydrate formation and potential blockage. As vaporized methanol flows along the pipeline from the injection point at the wellhead in Fig. 1, it dissolves into any produced or condensed water.

Hydrate blockages occur in the free water, usually just downstream from water accumulations where there is a change in flow geometry (e.g., a bend or pipeline dip along an ocean-floor depression), or some nucleation site (e.g., sand or weld slag). Hydrate inhibition occurs at the interface of the aqueous liquid, which contains most of the methanol, rather than in the bulk vapor, or oil/condensate. This example illustrates hydrate avoidance, attained by injecting sufficient methanol in the gas to partition 23 wt% methanol in the free water, thus preventing the flowline conditions from entering the hydrate thermodynamic-stability region. However, high water production requires large amounts of methanol to be injected, making the economics poor and sometimes impractical. Therefore, methods other than avoidance have been considered. Thermodynamics Predictions Successful avoidance is enabled by newer-generation hydrate programs for thermodynamic-formation-condition predictions with and without inhibitors. Fig. 2 shows the accuracy of five common hydrate-prediction software programs compared against published uninhibited-hydrateequilibrium data as of 2002 for single-, binary-, and ternary-hydrate-guest molecules, as well as natural gases, black oils, and gas condensates. Earlier programs provided only the initial limits of hydrate formation. New thermodynamic-program generation is based on hydrate-phase measurements and can predict useful additional conditions, such as hydrate-phase amounts. Fig. 2 shows that the average absolute error in temperature for the five prediction programs was within 1 K. On a pressure basis, predictions can be expected to be accurate to within 10%. These prediction errors approximate the errors of the experiments and are acceptable for engineering purposes. It may not be practical to improve predictions for the majority of cases. However, improved hydratethermodynamics measurements and predictions are needed, particularly for high (>50 wt%) concentrations of inhibitors and those mixtures. Thermodynamics predictions enable the flow-assurance engineer to avoid hydrate-formation conditions in a flow-

90

JPT DECEMBER 2009

Average Absolute Error in Temperature, K

0.75

CSMGem CSMHYD DBRHydrate Multiflash PVTsim

1.36

0.5

0.25

NA

0.25

Single (632 pts)

Binary (747 pts)

Ternary (89 pts)

Natural Gas (72 pts)

BO and GC (10 pts)

sH (135 pts)

Types of Hydrates

Fig. 2Absolute-temperature errors in common prediction simulators compared against all published hydrocarbon-hydrate-formation data (single, binary, and ternary refer to the number of guest molecules; BO=black oil; GC=gas condensate; sH=structure H).

line, but at a cost of insulation for high temperatures, or of inhibitor injectioncosts that may become unacceptable for marginally economic production projects.

of the hydrate shell, the particle remains the same size as the original water droplet. 3. Within each hydrate shell, shrinking-core dropletssimulators continue compared to grow inward, asall a published function of mass Fig. 2Absolute-temperature errors in common prediction against hydroHydrate-Risk Managementdata (single, binary, and ternary transfer of to the guest through the oil and the hydrate carbon-hydrate-formation refer the types of guest molecules; BO= black oil; =gasnumber condensate; sH=hydrate structure). In a GC growing of flow-assurance situations, hydrate- shell, and of heat transfer dissipating the energy from risk management is more economical than avoidance. One hydrate formation. Possible free water coating each dropaspect of hydrate-risk management is to allow hydrate par- let enables strong capillary attractive forces between the ticles to form, but to prevent hydrate-particle aggregation hydrated droplets. to a blockage by ensuring that the particles will flow, and 4. Hydrated droplets agglomerate, forming larger masses remain entrained in the oil phase. To move from avoidance that can plug the pipeline, as shown at the right in Fig. 3. to risk management, it is essential to quantify hydrate-forThis conceptual representation evolved from a decade of mation time dependence, on the basis of conceptual pictures both laboratory and flow-loop studies. In risk management, described below. as opposed to avoidance, the key to preventing hydrate-plug In hydrate kinetics, most of the initial quantitative kinetic formation is to prevent particle agglomeration by means of data came from the laboratory of Bishnoi and his colleagues shear, antiagglomerants, or other techniques, such as natuat the University of Calgary. Hydrate-kinetics experiments rally inhibited oils. One key implicit assumption is that all are difficult because hydrate formation is confounded by water is dispersed in the oil phase, an assumption that must overriding transient phenomena such as heat- and mass- be corrected in future developments. transfer effects. A strict separation of all three effects (i.e., forFrom the experiments to generate the conceptual repremation kinetics, mass transfer, and heat transfer) is required sentation in Fig. 3, six flow-assurance rules of thumb arise. for acceptable modeling. 1. The formation of emulsions/dispersions to keep the water/hydrates suspended in the oil phase is one key to preHydrate Blockages in Oil-Dominated Systems. Fig. 3 is a vent blockage formation. conceptual representation, with input from J.A. Abrahamson 2. Particle aggregation may be prevented by high shear (University of Canterbury, Christchurch, NZ), for hydrate stress. formation in an oil-dominated system with <50 vol% water 3. There are two requirements to prevent hydrate-plug cuts. formation in oil-dominated systems: Fig. 3 shows four steps in hydrate-plug formation along a (a) Hydrate particles must be in low concentration flowline length. (<50 vol%). 1. Water is entrained in an oil-continuous-phase emulsion (b) Particle aggregation is prevented by particles being oilas droplets, typically less than 50-m diameter, the result of wet through oil chemistry, or by application of antiagoil chemistry and shear. glomerant chemicals. 2. As the flowline enters the hydrate-formation region, 4. The closer the operating conditions are to the hydratehydrate grows rapidly (at approximately 1 mm/3 sec) at dissociation conditions, the stronger the interactions between the oil/water interface, forming thin (10- to 30-m-thick) hydrate particles, because of attractive capillary forces from a hydrate shells around the droplets. Even with the formation quasiliquid layer at the particle interface.

JPT DECEMBER 2009

NA

91

DISTINGUISHED AUTHOR SERIES


Water Entrainment Hydrate-Shell Growth

Agglomeration

Plug

Gas Oil Water


Time/Distance Capillary Attraction Hydrate Shells

Fig. 3Conceptual representation of hydrate formation in an oil-dominated system.

5. The formation and dissociation of hydrates can cause coalescence of water drops in water-in-oil systems. This coalesced free-water phase is prone to hydrate-blockage formation. 6. Like other deposits, freshly formed hydrates are more porous and malleable than are hydrates that have time to age and solidify. The aging process (something akin to Ostwald ripening) causes a more-dense crystal mass, making dissociation of the plug increasingly difficult. Cold Stabilized Flow. The above qualitative rules of thumb have many flow-assurance applications, and they are quantified in software programs that can predict approximately where and when hydrates can form in oil-dominated flowlines. However, it may be worthwhile to consider one patented application that may be field-tested in the near future. There are two main, patented, cold or stabilized flow concepts (Lund et. al. 2004; Talley et al. 2007). In each process, the key principle is to emulsify and convert free water to hydrate as entirely and rapidly as possible. Without a freewater phase to encourage hydrate-particle aggregation, as in Rules of Thumb 3 and 4 in the preceding subsection, hydrate particles will not aggregate but will flow with the oil phase, much like dry snow is difficult to compact/aggregate into a snowball. Conversely, without a significant gas phase, insufficient hydrate may form to plug a flowline.

Hydrate Blockages in Condensate Systems. Light condensate systems differ from oil-dominated systems because emulsified water droplets do not form without high shear because of low viscosity and the lack of surface-active components. There is a severe lack of published hydrate field data for condensate-flowline plugs, although the general narrowing of the flow path for hydrate formation has gained acceptance, as shown in Fig. 4. Three additional rules of thumb for hydrate formation from a condensate were determined by measurements in a liquid/condensate flow loop (Nicholas et al. 2009), in conjunction with several laboratory measurements of adhesive forces between condensate hydrates and pipe materials. 1. Hydrates formed in bulk condensate may not deposit on the wall in the absence of water-wet walls. 2. Hydrates formed at the pipe surface will remain on the wall. (a) High concentrations of dissolved water provide a uniform, dispersed deposit along the flowline. (b) Free water results in a localized, early deposit as the flowline enters the hydrate-stability region. 3. Hydrate deposits can be dissociated with or without chemicals by flowing an undersaturated condensate past the hydrate deposit, and by use of methanol dissolved in the hydrocarbon. Typically, methanol is injected into the flowline where it reaches equilibrium with the species present. With hydrate deposition on flowline walls, the mechanism for condensate-hydrate-plug formation may differ significantly from that of an oil-dominated-system plug. In condensate systems, sloughing and particle jamming likely will occur to form a plug. In oil-dominated systems, particle aggregation will increase the apparent viscosity for effective plugging. Sloughing and jamming are subjects of current research. The above rules of thumb have several important implications for the operation of a condensate flowline. For example, if a platform dehydrator operation has failed such that free water occurs, the export line should be shut for immediate remediation. However, if dissolved water is above the hydrate-equilibrium concentration, then corrective action may be taken by bringing the dehydrator to acceptable limits, for dissolution of the hydrate wall deposit. Future of Hydrates in Flow Assurance The above conceptual representation presents a beginning for understanding phenomena associated with hydrate-plug formation in relatively low-water-cut systems. The concepts provide suggestions for incorporation in transient multiphase-flow software to enable the flow-assurance engineer to determine the risk of hydrate-plug formation in an oildominated flowline. In addition, the concepts may be used to deal with systems that fail to meet design expectations, and that are found to be at risk. These conceptual representations are being extended to other cases in which a free-water phase exists in addition to emulsified water in oil. In the next decade, it may be possible for a flow-assurance engineer to determine the risk of flowline plugging resulting from hydrate formation by three steps. 1. Obtain an uncontaminated sample of an oil for a new field.

Fig. 4Hydrate formation narrowing the channel by wall deposition, (Courtesy of G. Hatton, formerly of Southwest Research Institute). Note: The bent tube at the top blows nitrogen against the window for visibility.

92

JPT DECEMBER 2009

Arctic sandstones under existing infrastructure (10s of Tcf in place) Arctic sandstones away from infrastructure (100s of Tcf in place) Deepwater sandstones (1,000s of Tcf in place) Nonsandstone marine reservoirs with permeability (unknown) Massive surficial and shallow nodular hydrate (unknown) Marine reservoirs with limited permeability (100,000s of Tcf in place)

Reserves (200 Tcf) Expected reserves growth (500 Tcf) Undiscovered (1,500 Tcf recoverable) Remaining unrecoverable (unknown)

Fig. 5Hydrate resource (left) relative to the conventional natural-gas resource (right) for the United States (Boswell and Collett 2006).

2. Perform bench-scale hydrate tests to determine emulsion characteristics with chemistry and shear over a representative temperature range. 3. Forecast plug-formation risk by use of the fluid information from Step 2 and transient multiphase computer software. With risk assessment in hand, the engineer can determine which prevention steps are required. Ultimately, such risk assessments must be validated against field data for hydrate formation. In Nature: Energy and Climate Change Two excellent reviews of hydrates in nature have been published by the National Petroleum Council (Kleinberg 2007) and by the Council of Canadian Academies (Grace et al. 2008). The enormous potential for hydrated energy is indicated by the worldwide amount of methane in hydrates, with estimates ranging from 0.9401017 scf, and may be almost two orders of magnitude greater than conventional reserves. In the US, the relative magnitudes of hydrate and conventional gas are shown in Fig. 5. Current estimates of the volume of natural gas trapped in hydrates within the US are on the order of 200,000 Tcf. Even if only a small fraction of this volume is recoverable, this natural-gas resource could provide an enormous contribution relative to the current US domestic consumption level (22 Tcf/yr) and expected future growth in demand. However, it is important to note two vital starting points. The total amount of methane in the ocean is uncertain, but estimated to be greater than that in the permafrost by a factor of 100, although permafrost hydrates may be more accessible and frequently have higher concentrations. It is not only the total amount of hydrate, but also the concentration and location of the resource that determines recoverability of methane from hydrates in nature. In Fig. 5

for example, the recovery of methane from the lower portion of the hydrate pyramid (with limited permeability) is problematic. Relative to the amount of conventional gas available, methane from hydrates is considered precommercial; therefore, industry/government partnerships are required for development. Countries with a high energy demand such as Japan, India, China, and South Korea are mounting large campaigns to develop hydrated energy; the goal for Japan is commercial productivity by 2015. Because of space limitations, summary statements are provided below with details that may be investigated in the cited references. Natural hydrates most likely are to be found in sandy sediments because silty sediments typically have lean hydrate accumulations (Grace et al. 2008). In the 2002 Mallik well, hydrates were determined to be pore-filling and provided some mechanical stability to the sediment. Depressurization is thought to be the most economical way to produce hydrates. The technology and economics of hydrates in Arctic permafrost and hydrates in the marine environment are sufficiently different and must be treated separately. Arctic-Permafrost Hydrates. There are no great technical deterrents to recovery of energy. Hydrates have been produced for short periods in the 2007 Mt. Elbert well (Boswell et al. 2008), and in the 6-day 2008 Mallik depressurization, which had average flow rates of 70 Mcf/D, with peak rates as high as 160 Mcf/D (Grace et al. 2008). Hydrates have, as the largest technical concern, wellbore/ reservoir geomechanical stability during production. Hydrates require continuous multiyear production testing to enable reservoir modelers to eliminate transient effects acceptably and to assess commercial feasibility (Kleinberg 2007).

JPT DECEMBER 2009

93

DISTINGUISHED AUTHOR SERIES


Hydrates provide an opportunity, during such production tests, for innovative technologies to be assessed such as CO2 displacement of CH4. Hydrates provide an acceptable place (ease of access, high concentrations at sweet spots) for developments that can be transferred to the ocean in the future. Hydrates may be recoverable economically, particularly in places where there is access to existing infrastructure (Walsh et al. 2009). Marine Gas Hydrates. These resources are less advanced developmentally than Arctic-permafrost hydrates because gas from marine hydrates has not yet been produced. These resources have a major technology challenge of a reliable method to find hydrates. (Kleinberg 2007). The common bottom-simulating reflector is a first-order-detection method, which frequently is unreliable. These resources require a multisite drilling expedition for reliable assessment and recovery, which would have a very high expense. The 2006, 113-day offshore Indian hydrate-exploratory expedition required USD 36 million. International cooperation is required to share expenses and results. These resources have unconventional hydrate-gas-recovery economics that are two to three times more expensive than conventional offshore gas, when existing infrastructure is unavailable for either (Walsh et al. 2009). Climate Change. While hydrate production is expected to have no unusual environmental concerns, methane evolution from natural hydrate deposits is considered here. The isotopic record supports global warming from hydratedmethane evolution approximately 600 million years ago. More recently, there is conflicting evidence from analysis of the isotopic record from the late Quaternary. It appears that hydrates may have been relatively stable for the last 10,000 years (Grace et al. 2008). Little is known about methane evolution from hydrates in nature. Methane may be oxidized before reaching the upper atmosphere. The most-active locations for methane evolution likely are in the marine permafrost. Hydrated-methane evolution is not expected to be a major environmental factor in this century. Any methane evolution is likely to be chronic, rather than abrupt. Conclusion For the last three-quarters of a century, hydrates have been a major flow-assurance concern to the oil and gas industry. Progress in quantifying hydrate formation provides the engineer with new flow-assurance tools, both in oil-dominated and condensate-dominated systems. In the future, it may be possible to predict the hydrate-plug risk of an oil-production system from a few simple measurements, aided by transient multiphase software, to indicate emulsion stability and jamming of hydrated particles. Currently, natural-gas hydrates are precommercial and require government/industry partnerships to prove viability. With increasing international cooperation, the next decade will witness methane production from hydrates as a member of the energy-resource spectrum. The climate change from hydrates likely is chronic, rather than abrupt. Dedication This article is dedicated to Yuri F . Makogon, on the occasion of his retirement after a career of hydrate development. References
Boswell, R., Amato, R., Coffin, R., Collett, T., Dellagiarino, G., Fisk, R., Gettrust, J. et al. 2006. An Interagency Roadmap for Methane Hydrate Research and Development. US DOE Office of Fossil Energy (July 2006), http://www.fe.doe.gov/ programs/oilgas/publications/methane_hydrates/mh_interagency_plan.pdf. Downloaded 22 September 2009. Boswell, R. and Collett, T. 2006. The Gas Hydrates Resource Pyramid. Fire In The Ice (The National Energy Technology Laboratory Methane Hydrate Newsletter) Fall 2006: 57. Boswell, R., Hunter, R., Collett, T., Digert, S., Hancock, S., and Weeks, B. 2008. Investigation of Gas Hydrate Bearing Sandstone Reservoirs at the Mount Elbert Stratigraphic Test Well, Milne Point, Alaska. Proceedings of the 6th International Conference on Gas Hydrates, Vancouver, Canada. 610 July. Grace, J., Collett, T., Colwell, F ., Englezos, P., Jones, E., Mansell, R., Meekison, J.P. et al. 2008. Energy From Gas Hydrates: Assessing the Challenges and Opportunities for Canada. Expert Panel on Gas Hydrates Project Report, Council of Canadian Academies, Ottawa, Canada (September 2008), http://www.scienceadvice.ca/documents/(2008-11-05)%20 Report%20on%20GH.pdf. Downloaded 22 September 2009. Kleinberg, R.L. 2007. Topic Paper 24: Hydrates. In Facing the Hard Truths about Energy, National Petroleum Council Report, Washington, DC (18 July 2007), http://www. npchardtruthsreport.org/. Downloaded 22 September 2009. Lund, A., Lysne, D., Larson, R., and Hjarbo, K.W. 2004. Method and system for transporting a flow of fluid hydrocarbons containing water. US Patent No. 6,774,276; International (PCT) Patent No. WO/2000/025062; Norwegian Patent No. NO 311,854. Nicholas, J.W., Koh, C.A., and Sloan, E.D. 2009. A preliminary approach to modeling gas hydrate/ice deposition in a liquid condensate system. AIChE Journal 55 (7): 1889 1897. doi: 10.1002/aic.11921. Notz, P.K. 1994. Discussion of the Paper The Study of the Separation of Nitrogen from Methane by Hydrate Formation Using a Novel Apparatus. Annals of the New York Academy of Sciences 715 (1): 425429. doi: 10.1111/ j.1749-6632.1994.tb38855.x. Sloan, E.D. Jr. and Koh, C.A. 2008. Clathrate Hydrates of Natural Gases, third edition, Vol. 119. Boca Raton, Florida: Chemical Industries, CRC Press. Talley, L.D., Turner, D.J., and Priedeman, D.K. 2007. Method of generating a non-plugging hydrate slurry. International (PCT) Patent No. WO/2007/095399. Walsh, M.R., Hancock, S.H., Wilson, S.J., Patil, S.L., Moridis, G.J., Boswell, R., Collett, T.S., Koh, C.A., and Sloan, E.D. 2009. Preliminary report on the commercial viability of gas production from natural gas hydrates. Energy Economics JPT 31 (5): 815823. doi:10.1016/j.eneco.03.006.

94

JPT DECEMBER 2009

Вам также может понравиться