Hostname: page-component-848d4c4894-8bljj Total loading time: 0 Render date: 2024-06-14T05:08:41.309Z Has data issue: false hasContentIssue false

Investigation of the Internal Structure of a Modern Seafloor Hydrothermal Chimney With a Combination of EBSD, EPMA, and XRD

Published online by Cambridge University Press:  20 May 2020

Matthew Glenn*
Affiliation:
Microbeam Laboratory, CSIRO Mineral Resources, Clayton, VIC, Australia
Siyu Hu
Affiliation:
CSIRO Mineral Resources, Kensington, WA, Australia
Stephen Barnes
Affiliation:
CSIRO Mineral Resources, Kensington, WA, Australia
Aaron Torpy
Affiliation:
Microbeam Laboratory, CSIRO Mineral Resources, Clayton, VIC, Australia
Anthony E. Hughes
Affiliation:
Microbeam Laboratory, CSIRO Mineral Resources, Clayton, VIC, Australia
Colin M. MacRae
Affiliation:
Microbeam Laboratory, CSIRO Mineral Resources, Clayton, VIC, Australia
Nathan A.S. Webster
Affiliation:
Microbeam Laboratory, CSIRO Mineral Resources, Clayton, VIC, Australia
Nicholas C. Wilson
Affiliation:
Microbeam Laboratory, CSIRO Mineral Resources, Clayton, VIC, Australia
Joanna Parr
Affiliation:
CSIRO Mineral Resources, North Ryde, NSW, Australia
Ray Binns
Affiliation:
CSIRO Mineral Resources, North Ryde, NSW, Australia
*
*Author for correspondence: A.M. Glenn, E-mail: matthew.glenn@csiro.au
Get access

Abstract

Samples from the sphalerite-dominated zone of a seafloor massive sulfide chimney, the Satanic Mills Chimney of the PACMANUS hydrothermal field, have been investigated to determine the internal macrostructure and microstructure of this zone, the phases present, and the distribution of metals. A combination of electron probe microanalysis, electron backscattered diffraction, and x-ray diffraction has been used. At the macroscale, this zone of the chimney wall is heavily porous and is comprised primarily of sphalerite, enclosing minor chalcopyrite, pyrite, and wurtzite. A PbAs sulfosalt layer of possible microbial origins is present at the outer edge of the sphalerite matrix, next to a pore. The sphalerite has grown in globules on the order of 300 μm in diameter. At the microscale, the sphalerite features a colloform texture and a duplex-type grain structure consisting of either fine-grain regions in the center surrounded by coarse-grained regions or radiating coarse grains only. Pb- and As-rich bands have been detected in the colloform sphalerite, and growth twins have been observed in both the sphalerite and chalcopyrite crystals. A qualitative description of the growth of a typical globule is given, including nucleation, crystal growth, and solute redistribution.

Type
Australian Microbeam Analysis Society Special Section AMAS XV 2019
Copyright
Copyright © Microscopy Society of America 2020

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Auzende, J-M & Urabe, T (1996 a). Cruise explores hydrothermal vents of the Manus Basin. EOS Trans 77(26), 244245.CrossRefGoogle Scholar
Auzende, J-M, Urabe, T, Ruellan, E, Charbroux, D, Charlou, J-L, Gena, K, Gamo, T, Henry, K, Matsubayashi, O, Matsumoto, T, Naka, J, Nagaya, Y & Okamura, K (1996 b). “Shinkai 6500” dives in the Manus Basin: New starmer Japanese-French program. JAMSTEC J Deep Sea Res 12, 323334.Google Scholar
Barrie, CD, Boyce, AJ, Boyle, AP, Williams, PJ, Blake, K, Wilkinson, JJ, Lowther, M, McDermott, P & Prior, DJ (2009). On the growth of colloform textures: A case study of sphalerite from the Galmoy Ore Body, Ireland. J Geol Soc Londn 166, 563582.CrossRefGoogle Scholar
Barriga, FJAS, Binns, RA, Miller, DJ & Herzig, PM (eds) (2007). Anatomy of an active felsic-hosted hydrothermal System. In Proceedings Ocean Drilling Program, Scientific Results 193. College Station, TX. doi:10.2973/odp.proc.sr.193.2007.CrossRefGoogle Scholar
Berkenbosch, A, de Ronde, CEJ, Gemmell, JB, McNeill, AW & Goemann, K (2012). Mineralogy and formation of black smoker chimneys from brothers submarine Volcano, Kermadec Arc. Econ Geol 107, 16131633.CrossRefGoogle Scholar
Binns, R (1995). PACMANUS: An active seafloor hydrothermal field on siliceous volcanic rocks in the eastern Manus Basin, Papua New Guinea. In Proceedings of The 1995 Pacrim Congress, Auckland, New Zealand, 19–22 November 1995, Mauk JD (Ed.), pp. 49–54, The Australasian Institute of Mining and Metallurgy.Google Scholar
Binns, RA, Parr, JM, Waters, JC, Gemmell, JB & Scott, SD (1999). Hydrothermal Activity at the PACMANUS Hydrothermal Field, Eastern Manus Basin, Papua New Guinea, Australia. Research Review, CSIRO Exploration and Mining, pp. 4–5.Google Scholar
Binns, RA & Scott, SD (1993). Actively forming polymetallic sulfide deposits associated with felsic volcanic rocks in the eastern Manus back-arc basin, Papua New Guinea. Econ Geol 88, 22262236.CrossRefGoogle Scholar
Boschen, RE, Rowden, AA, Clark, MR & Gardner, JPA (2013). Mining of deep-sea seafloor massive sulfides: A review of the deposits, their benthic communities, impacts from mining, regulatory frameworks and management strategies. Ocean Coastal Manag 84, 5467.CrossRefGoogle Scholar
Buerger, MJ (1945). The genisis of twin crystals. Am Mineral 30(7-8), 469482.Google Scholar
Butler, IB & Nesbitt, RW (1999). Trace element distributions in the chalcopyrite wall of a black smoker chimney: Insights from laser ablation inductively coupled plasma mass spectrometry (LA–ICP–MS). Earth Planet Sci Lett 167(3), 335345.CrossRefGoogle Scholar
Castaing, R (1952). Application des Sondes Electroniques a une methode d'analyse ponctuelle chimique et cristallographique. Thesis, Universite de Paris, Paris, France.Google Scholar
Corliss, JB, Dymond, J, Gordon, LI, Edmond, JM, Von Herzen, RP, Ballard, RD, Green, K, Williams, D, Bainbridge, A, Crane, K & Van Andel, TH (1979). Submarine thermal springs on the Galapagos Rift. Science 203(4385), 10731083.CrossRefGoogle ScholarPubMed
de Ronde, C, Faure, K, Bray, C, Chappell, D & Wright, I (2003). Hydrothermal fluids associated with seafloor mineralization at two southern Kermadec arc volcanoes, offshore New Zealand. Miner Deposita 38(2), 217233.CrossRefGoogle Scholar
de Ronde, CEJ, Hannington, MD, Stoffers, P, Wright, IC, Ditchburn, RG, Reyes, AG, Baker, ET, Massoth, GJ, Lupton, JE, Walker, SL, Greene, RR, Soong, CWR, Ishibashi, J, Lebon, GT, Bray, CJ & Resing, JA (2005). Evolution of a submarine magmatic-hydrothermal system; Brothers Volcano, southern Kermadec Arc, New Zealand. Econ Geol 100(6), 10971133.CrossRefGoogle Scholar
Flemings, MC (1974). Solidification Processing. New York, USA: McGraw Hill.CrossRefGoogle Scholar
Galley, AG, Hannington, MD & Jonasson, IR (2007). Volcanogenic massive sulphide deposits. In Mineral Deposits of Canada: A Synthesis of Major Deposit-Types, District Metallogeny, the Evolution of Geological Provinces, and Exploration Methods. Goodfellow, WD (Ed.). St. John's, NL, Canada: Geological Association of Canada, Mineral Deposits Division, Special Publication No. 5, pp. 141161.Google Scholar
Gena, K (2013). Deep sea mining of submarine hydrothermal deposits and its possible environmental impact in Manus Basin, Papua New Guinea. Procedia Earth Planet Sci 6(C), 226233.CrossRefGoogle Scholar
German, CR & Seyfried, WE (2014). Hydrothermal processes. In Treatise on Geochemistry, Holland HD & Turekian KK (Eds.), 2nd ed. Amsterdam, Netherlands: Elsevier Ltd., Ch 8.7, pp 191233.CrossRefGoogle Scholar
Goldfarb, MS, Converse, DR, Holland, HD & Edmond, JM (1983). The genesis of hot spring deposits on the East Pacific Rise, 21 degrees N. Econ Geol Mon 5, 184197.Google Scholar
Hannington, MD, Jonasson, IR, Herzig, PM & Petersen, S (1995). Physical and chemical processes of seafloor mineralization at mid-ocean ridges. In Seafloor Hydrothermal Systems, Physical, Chemical, Biological and Geological Interactions, Humphris, SE, ZierenBerg, RA, Mullineaux, LS, Thomson, RE (Eds.), pp. 115157. Washington, DC, USA: American Geophysical Union.Google Scholar
Harrowfield, IR, MacRae, CM & Wilson, NC (1993). Chemical imaging in electron microprobes. Proceedings of the 27th Annual MAS Meeting, New York, Microbeam Analysis Society, pp. 547–548.Google Scholar
Haymon, RM (1983). Growth history of hydrothermal black smoker chimneys. Nature 301(5902), 695.CrossRefGoogle Scholar
Haymon, RM, Fornari, DJ, Edwards, MH, Carbotte, S, Wright, D & Macdonald, KC (1991). Hydrothermal vent distribution along the East Pacific Rise crest (9°09′–54′N) and its relationship to magmatic and tectonic processes on fast-spreading mid-ocean ridges. Earth Planet Sci Lett 104(2), 513534.CrossRefGoogle Scholar
Herzig, PM & Hannington, MD (1995). Polymetallic massive sulfides at the modern seafloor a review. Ore Geol Rev 10(2), 95115.CrossRefGoogle Scholar
Hu, SY, Barnes, SJ, Matthew Glenn, A, Pagès, A, Parr, J, MacRae, C & Binns, R (2019). Growth history of sphalerite in a modern sea floor hydrothermal chimney revealed by electron backscattered diffraction. Econ Geol 114(1), 165176.CrossRefGoogle Scholar
Hu, SY, Barnes, SJ, Pagès, A, Parr, J, Binns, R, Verrall, M, Quadir, Z, Rickard, WDA, Liu, W, Fougerouse, D, Grice, K, Schoneveld, L, Ryan, C & Paterson, D (2020). Life on the edge: Microbial biomineralization in an arsenic- and lead-rich deep-sea hydrothermal vent. Chem Geol 533, 119438.CrossRefGoogle Scholar
Hu, SY, Evans, K, Fisher, L, Rempel, K, Craw, D, Evans, NJ, Cumberland, S, Robert, A & Grice, K (2016). Associations between sulphides, carbonaceous material, gold and other trace elements in polyframboids: Implications for the source of orogenic gold deposits, Otago Schist, New Zealand. Geochim Cosmochim Acta 180, 197213.CrossRefGoogle Scholar
Hu, SY, Pagès, A, Barnes, S, Parr, J, Binns, R & Paterson, D (2017). Synchrotron X-ray fluorescence microscopy study of the seafloor hydrothermal chimneys and associated mineralised microbes. Poster for Australian Synchrotron Symposium. Melbourne, Australia: Australian Synchrotron.Google Scholar
Humphris, SE, Herzig, PM, Miller, DJ, Alt, JC, Becker, K, Brown, D, Brügmann, G, Chiba, H, Fouquet, Y, Gemmell, JB, Guerin, G, Hannington, MD, Holm, NG, Honnorez, JJ, Iturrino, GJ, Knott, R, Ludwig, R, Nakamura, K, Petersen, S, Reysenbach, AL, Rona, PA, Smith, S, Sturz, AA, Tivey, K & Zhao, X (1995). The internal structure of an active sea-floor massive sulphide deposit. Nature 377(6551), 713.CrossRefGoogle Scholar
Keith, M, Häckel, F, Haase, KM, Schwarz-Schampera, U & Klemd, R (2016). Trace element systematics of pyrite from submarine hydrothermal vents. Ore Geol Rev 72(1), 728745.CrossRefGoogle Scholar
Kelley, D, Baross, J & Delaney, JR (2002). Volcanoes, fluids, and life at mid-ocean ridge spreading centers. Annu Rev Earth Planet Sci 30, 385491.CrossRefGoogle Scholar
Koski, RA, Clague, DA & Oudin, E (1984). Mineralogy and chemistry of massive sulfide deposits from the Juan de Fuca Ridge. Geol Soc Am Bull 95(8), 930945.2.0.CO;2>CrossRefGoogle Scholar
Koski, R, Jonasson, IR, Kadko, DC, Smith, V & Wong, F (1994). Compositions, growth mechanisms, and temporal relations of hydrothermal sulfide-sulfate-silica chimneys at The Northern Cleft Segment, Juan-De-Fuca Ridge. J Geophys Res-Solid Earth 99(B3), 48134832.CrossRefGoogle Scholar
Kristall, B, Kelley, DS, Hannington, MD & Delaney, JR (2006). Growth history of a diffusely venting sulfide structure from the Juan de Fuca Ridge: A petrological and geochemical study. Geochem Geophys Geosyst 7, 7.CrossRefGoogle Scholar
Kristall, B, Nielsen, D, Hannington, MD, Kelley, DS & Delaney, JR (2011). Chemical microenvironments within sulfide structures from the Mothra Hydrothermal Field: Evidence from high-resolution zoning of trace elements. Chem Geol 290(1), 1230.CrossRefGoogle Scholar
Kurz, W & Fisher, DJ (1998). “Fundamentals of Solidification”, Trans. Tech Publications, Switzerland.Google Scholar
Maslennikov, VV, Maslennikova, SP, Large, RR, Danyushevsky, LV, Herrington, RJ, Ayupova, NR, Zaykov, VV, Lein, AY, Tseluyko, AS, Melekestseva, IY & Tessalina, SG (2017). Chimneys in Paleozoic massive sulfide mounds of the Urals VMS deposits: Mineral and trace element comparison with modern black, grey, white and clear smokers. Ore Geol Rev 85, 64106.CrossRefGoogle Scholar
Mullin, JW (2001). Crystallization. Oxford, UK: Butterworth Heinemann.Google Scholar
Nowell, MM & Wright, SI (2004). Phase differentiation via combined EBSD and XEDS. J Microsc 213(3), 296305.CrossRefGoogle ScholarPubMed
Oudin, E & Constantinou, G (1984). Black smoker chimney fragments in Cyprus sulphide deposits. Nature 308(5957), 349.CrossRefGoogle Scholar
Parr, JM (1994). The preservation of pre-metamorphic colloform banding in pyrite from the broken Hill-Type Pinnacles Deposit, New South Wales, Australia. Miner Mag 58(9), 461471.CrossRefGoogle Scholar
Perfit, MR, Ridley, WI, Jonasson, IR, Barrie, CT & Hannington, MD (1999). Geologic, petrologic, and geochemical relationships between magmatism and massive sulfide mineralization along the eastern Galapagos spreading center. Rev Econ Geol 8, 75100. Volcanic Associated Massive Sulphide Deposits: Processes and Examples in Modern and Ancient Settings, Chapter 4.Google Scholar
Porter, DA & Easterling, KE (1981). Phase Transformations in Metals and Alloys. UK: Van Nostrand Reinhold, International.Google Scholar
Pouchou, J-L (1993). X-Ray Microanalysis of Stratified Specimens. Anal Chim Acta 283, 8197.CrossRefGoogle Scholar
Pouchou, J-L & Pichoir, F (1991). “Quantitative analysis of homogeneous or stratified microvolumes applying the model ‘PAP’”. In Electron Probe Quantitation, Heinrich, KFJ & Newbury, DE (Eds.), New York: Plenum Press, p. 31.CrossRefGoogle Scholar
Revan, MK, Genç, Y, Maslennikov, VV, Maslennikova, SP, Large, RR & Danyushevsky, LV (2014). Mineralogy and trace-element geochemistry of sulfide minerals in hydrothermal chimneys from the Upper-Cretaceous VMS deposits of the eastern Pontide orogenic belt (NE Turkey). Ore Geol Rev 63, 129149.CrossRefGoogle Scholar
Rona, PA, Klinkhammer, G, Nelsen, TA, Trefry, JH & Elderfield, H (1986). Black smokers, massive sulphides and vent biota at the Mid-Atlantic Ridge. Nature 321(6065), 33.CrossRefGoogle Scholar
Seewald, JS, Reeves, EP, Bach, W, Saccocia, PJ, Craddock, PR, Shanks, WC, Sylva, SP, Pichler, T, Rosner, M & Walsh, E (2015). Submarine venting of magmatic volatiles in the Eastern Manus Basin, Papua New Guinea. Geochim Cosmochim Acta 163, 178199.CrossRefGoogle Scholar
Shanks, WCP & Thurston, R (2012). Volcanogenic massive sulfide occurrence model. Scientific Investigations Report 2010-5070-C, U.S. Geological Survey, Reston, VA, USA.Google Scholar
Shannon, RD (1976). Revised effective ionic radii and systematic studies of interatomic distances in Halides and Chalcogenides. Acta Cryst A32, 751767.Google Scholar
Tao, C, Lin, J, Guo, S, Chen, YJ, Wu, G, Han, X, German, CR, Yoerger, DR, Zhou, N, Li, H, Su, X & Zhu, J (2012). First active hydrothermal vents on an ultraslow spreading center; Southwest Indian Ridge. Geology 40(1), 4750.CrossRefGoogle Scholar
Tivey, M (2007). Generation of Seafloor Hydrothermal Vent Fluids and Associated Mineral Deposits. Oceanography 20(1), 5065.CrossRefGoogle Scholar
Tivey, MK (1995). The influence of hydrothermal fluid composition and advection rates on black smoker chimney mineralogy: Insights from modeling transport and reaction. Geochim Cosmochim Acta 59(10), 19331949.CrossRefGoogle Scholar
Tivey, MK, Laflamme, BD & McDuff, RE (1986). Mineral precipitation in “black smoker” chimney walls; a 1-dimensional model of heat and mass transport. Trans Am Geophys Union 67(44), 1283.Google Scholar
Tivey, MK & McDuff, RE (1990). Mineral precipitation in the walls of black smoker chimneys: A quantitative model of transport and chemical reaction. J Geophys Res: Solid Earth 95(B8), 1261712637.CrossRefGoogle Scholar
Tivey, MK, Stakes, DS, Cook, TL, Hannington, MD & Petersen, S (1999). A model for growth of steep-sided vent structures on the Endeavour Segment of the Juan de Fuca Ridge: Results of a petrologic and geochemical study. J Geophys Res: Solid Earth 104(B10), 2285922883.Google Scholar
Tregoning, P (2002). Plate kinematics in the western Pacific derived from geodetic observations. J Geophys Res: Solid Earth 107(B1), 18. ECV 7CrossRefGoogle Scholar
Von Damm, KL (1995). Controls on the chemistry and temporal variability of seafloor hydrothermal fluids. In Seafloor Hydrothermal Systems, Physical, Chemical, Biological and Geological Interactions, Humphris, SE, ZierenBerg, RA, Mullineaux, LS, Thomson, RE (Eds.),, pp. 222247. Washington, DC, USA: American Geophysical Union.Google Scholar
Wilson, N & MacRae, C (2005). An automated hybrid clustering technique applied to spectral data sets. Microsc Microanal 11(S02), 434435.CrossRefGoogle Scholar
Wilson, N, MacRae, C & Torpy, A (2008). Analysis of combined multi-signal hyperspectral datasets using a clustering algorithm and visualisation tools. Microsc Microanal 14(2), 764765.CrossRefGoogle Scholar
Wohlgemuth-Ueberwasser, CC, Viljoen, F, Petersen, S & Vorster, C (2015). Distribution and solubility limits of trace elements in hydrothermal black smoker sulfides: An in-situ LA-ICP-MS study. Geochim Cosmochim Acta 159, 1641.CrossRefGoogle Scholar
Yeats, CJ, Parr, JM, Binns, RA, Gemmell, JB & Scott, SD (2014). The SuSu Knolls hydrothermal field, eastern Manus Basin, Papua New Guinea: An active submarine high-sulfidation copper-gold system. Econ Geol Bull Soc Econ Geol 109(8), 22072226.CrossRefGoogle Scholar
Yuan, B, Yu, H, Yang, Y, Zhao, Y, Yang, J, Xu, Y, Lin, Z & Tang, X (2018). Zone refinement related to the mineralization process as evidenced by mineralogy and element geochemistry in a chimney fragment from the Southwest Indian Ridge at 49.6°E. Chem Geol 482, 4660.CrossRefGoogle Scholar
Zhao, J, Liang, J, Long, X, Li, J, Xiang, Q & Zhang, J (2018). Genesis and evolution of Framboidal Pyrite and its Implications for the ore-forming process of Carlin-Style Gold Deposits, Southwestern China. Ore Geol Rev 102, 426436.CrossRefGoogle Scholar