Hostname: page-component-8448b6f56d-42gr6 Total loading time: 0 Render date: 2024-04-23T14:42:21.029Z Has data issue: false hasContentIssue false

ASSERTIONS OF ENTITLEMENT TO THE OUTER CONTINENTAL SHELF IN THE CENTRAL ARCTIC OCEAN

Published online by Cambridge University Press:  27 February 2017

Bjørn Kunoy*
Affiliation:
Ministry of Foreign Affairs and Trade of the Faroe Islands, bjornk@uvmr.fo.

Abstract

The legal and technical issues relating to the outer continental shelf entitlements in the Central Arctic Ocean present several challenges, most of which are to be resolved in accordance with Article 76 of the United Nations Convention on the Law of the Sea. Recently, two coastal States in the Central Arctic Ocean have made fully fledged submissions relating to the Arctic to the Commission on the Limits of the Continental Shelf. Russia has made a revised submission that is currently being considered by the Commission on the Limits of the Continental Shelf. The submission of Denmark/Greenland will most likely only be considered in 10 or 15 years time.

Type
Articles
Copyright
Copyright © British Institute of International and Comparative Law 2017 

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.)

Footnotes

The views expressed in this article are strictly personal to the author. The author would like to thank Michael A Becker and an anonymous reviewer for their comments to a previous version of this article. The author is, of course, solely responsible for all of the observations and judgments contained in this article.

References

1 1833 UNTS, 396 (entered into force on 16 November 1994).

2 Art 76(1) of UNCLOS provides: ‘The continental shelf of a coastal State comprises the seabed and subsoil of the submarine areas that extend beyond its territorial sea throughout the natural prolongation of its land territory to the outer edge of the continental margin, or to a distance of 200 nautical miles from the baselines from which the breadth of the territorial sea is measured where the outer edge of the continental margin does not extend up to that distance.’

3 Excerpt from art 4 of Annex II to UNCLOS.

4 Russia lodged an initial submission on 20 December 2001, which has been supplemented by two revised partial submissions, the latter of which covers areas in the Central Arctic Ocean and was submitted to the CLCS on 5 August 2015. Norway lodged a partial submission on 27 November 2006, which covers a relatively small area to the north of the Svalbard archipelago. On 30 March 2009 the CLCS finalized its considerations of the partial submission and made corresponding recommendations to Norway. On 15 December 2014 Denmark/Greenland submitted a partial submission which covers a relatively large area of the Central Arctic Ocean.

5 See Preliminary Information concerning the outer limits of the continental shelf of Canada in the Arctic Ocean, which Canada pursuant to SPLOS/Decision 183 submitted to the CLCS on 6 December 2013. <http://www.un.org/depts/los/clcs_new/submissions_files/preliminary/can_pi_en.pdf>.

6 The United States of America is not a State Party to UNCLOS but is undertaking the collection of data and information with a view to document claimed entitlement to the areas of interest. See <http://www.continentalshelf.gov/>. This article will not discuss the question whether art 76 reflects customary international law. Yet, it should be noted that the ICJ has stated that ‘[t]he Court considers that the definition of the continental shelf set out in Article 76, paragraph 1, of UNCLOS forms part of customary international law’. ICJ, Territorial and Maritime Dispute (Nicaragua v Colombia), Judgment of 18 November 2012, para 118.

7 ITLOS, Dispute concerning delimitation of the maritime boundary between Bangladesh and Myanmar in the Bay of Bengal (Bangladesh/Myanmar), Judgment of 15 March 2012, para 438.

8 The proposed outer limits of Denmark/Greenland relating to the Central Arctic Ocean, as these appear in the publicly available data, demonstrate that parts of the Gakkel Ridge is within the proposed outer limits. See <http://www.un.org/depts/los/clcs_new/submissions_files/dnk76_14/dnk2014_es.pdf>.

9 R Macnab et al. argue that ‘[o]nly two regions appear to be exempt from this projected jurisdiction: a small area in the Mendeleev Abyssal Plain, and a larger one that encompasses the Gakkel Ridge, an oceanic spreading centre. These will remain a part of the Area, with resources that fall within the jurisdiction of the International Seabed Authority.’ R Macnab et al., ‘Cooperative Preparations for Determining the Outer Limit of the Juridical Continental Shelf in the Arctic Ocean: A Model for Regional Collaboration in Other Parts of the World?’ (Spring 2001) IBRU Boundary and Security Bulletin 94.

10 See Recommendations of the Commission on the Limits of the Continental Shelf in regard to the submission made by the Cook Islands in respect of the Manihiki Plateau on 16 April 2009, adopted on 19 August 2016, <http://www.un.org/depts/los/clcs_new/submissions_files/cok23_09/2019_08_19_COM_REC_COK.pdf >.

11 Art 76(4)(b) of UNCLOS provides ‘[i]n the absence of evidence to the contrary, the foot of the continental slope shall be determined as the point of maximum change in the gradient at its base’.

12 Excerpt from art 76(4)(b) of UNCLOS.

13 See point 6.1.9 of the Scientific & Technical Guidelines of the Commission on the Limits of the Continental Shelf (Guidelines) in which it is provided that where the evidence to the contrary rule is applicable, the notions ‘natural prolongation’ and ‘submerged prolongation’ in paras 1 and 3 of art 76, respectively ‘clarify concepts such as natural prolongation of the land territory to the outer edge of the continental margin in the geological sense of these terms, which require the consideration of tectonics, sedimentology and other aspects of geology’.

14 In its recommendations to Argentina, the CLCS observed that ‘regarding the criteria to be applied for the establishment of a foot of the continental slope based on evidence to the contrary: the base and foot of the continental slope should not be located seaward of the region where the [seaward dipping reflectors] sequence terminates; the base and the foot of the continental slope should not be located seaward of the region where the thickness of the crust reduces to typical oceanic crustal values further seaward; and the specific seaward dipping reflector chosen as the “last unequivocally identifiable seaward dipping reflector” at the end of of the [seaward dipping reflector] sequence should be of sufficient coherency and impedience.’ Summary of Recommendations of the Commission on the Limits of the Continental Shelf in regard to the submission made by Argentina on 21 April 2009, adopted by the CLCS on 11 March 2016, para 49. <http://www.un.org/depts/los/clcs_new/submissions_files/arg25_09/2016_03_11_COM_SUMREC_ARG.pdf>.

15 ibid.

16 Recommendations of the Commission on the Limits of the Continental Shelf in regard to the submission made by the Cook Islands in respect of the Manihiki Plateau on 16 April 2009, para 58.

17 Art 76(5) of UNCLOS provides: ‘The fixed points comprising the line of the outer limits of the continental shelf on the seabed, drawn in accordance with paragraph 4 (a)(i) and (ii), either shall not exceed 350 nautical miles from the baselines from which the breadth of the territorial sea is measured or shall not exceed 100 nautical miles from the 2,500 metre isobath, which is a line connecting the depth of 2,500 metres.’ On this issue see Kunoy, B, Heinesen, MV and Mørk, F, ‘Appraisal of Applicable 2,500 m Depth Constraint Lines for the Purposes of the Establishment of the Outer Limits of the Continental Shelf’ (2010) 41 ODIL 357–79CrossRefGoogle Scholar.

18 The 100 nm distance line measured from the applicable 2500 metre isobath is referred to as the ‘depth constraint’.

19 See the recommendations of the CLCS to Australia in which it refused to classify the Williams Ridge (WR) as a submarine elevation that is a natural component of the continental margin. The CLCS stressed that ‘the data submitted for the WR seems to give only indirect evidence of its nature and origin and the Commission is of the opinion that the geological origin of the WR still remains unresolved. The Commission therefore questions whether the application of paragraph 7.3.1(b) of the Guidelines is justified in the case of the WR. Therefore the Commission does not consider it justified that the WR is regarded a submarine elevation that is a natural component of the continental margin in the sense of article 76, paragraph 6’. Recommendations of the Commission on the Limits of the Continental Shelf in regard to the Submission made by Australia on 15 November 2004, adopted on 9 April 2008, para 51 <http://www.un.org/depts/los/clcs_new/submissions_files/aus04/Aus_Recommendations_FINAL.pdf>.

20 The first sentence of art 76(6) of UNCLOS provides: ‘On submarine ridges, the outer limit of the continental shelf shall not exceed 350 [M] from the baselines.’

21 The proposed outer limits of Denmark/Greenland, as these appear in the publicly available data, demonstrate that the claim of outer continental shelf from Greenland extends up to approximately 950 M from the baselines. See <http://www.un.org/depts/los/clcs_new/submissions_files/dnk76_14/dnk2014_es.pdf>.

22 Coastal States appear vested with a right to block the CLCS from considering a submission from another State with which it has overlapping claims of entitlement. Rule 5(a) of Annex I to the Rules of Procedure of the CLCS provides that ‘[i]n cases where a land or maritime dispute exists, the Commission shall not consider and qualify a submission made by any of the States concerned in the dispute. However, the Commission may consider one or more submissions in the areas under dispute with prior consent given by all States that are parties to such a dispute.’ All the relevant coastal States in the Central Arctic Ocean have notified the United Nations Secretary General that they do not object to the CLCS considering the partial revised submission of Russia of August 3, 2015 (Denmark, United States of America and Canada submitted their respective notes on 7 October 2015, 30 October 2015 and on 30 November 2015; see <http://www.un.org/depts/los/clcs_new/submissions_files/submission_rus_rev1.htm>. Nor do they object to the CLCS considering the submission of Denmark/Greenland of December 15, 2014 (Norway, Canada, the Russian Federation and the United States of America submitted their respective notes on 17 December 2014, 29 December 2014, 21 July 2015 and 30 October 2015; see <http://www.un.org/depts/los/clcs_new/submissions_files/submission_dnk_76_2014.htm>. The entitlement to outer continental shelf stemming from Norway has already been considered by the CLCS. The CLCS approved the claimed entitlement of Norway stemming from the Yermak Plateau.

23 According to Rule 51(4)(ter) of the Rules of Procedure of the Commission on the Limits of the Continental Shelf (Rules of Procedure) ‘[t]he submissions shall be queued in the order they are received’. CLCS/40/Rev.1. The submission of Denmark/Greenland is number 76 in the queue. The most recent establishment of a sub-commission relates to the consideration of the submission of the Seychelles concerning the Northern Plateau Region, which is number 39 in the queue.

24 Notwithstanding the rule that submissions shall be considered in the order they are received, revised submissions shall be ‘considered on a priority basis notwithstanding the queue’. Statement by the Chairperson of the Commission on the progress of work in the Commission (CLCS/68) of 17 September 2010, para 57.

25 According to art 5 of Annex II to UNCLOS, ‘[u]nless the Commission decides otherwise, the Commission shall function by way of sub-commissions composed of seven members, appointed in a balanced manner taking into account the specific elements of each submission by a coastal State’. The Members of the sub-commission established to consider the partial revised submission of Russia are LF Awosika, G Carrera (Chair), MB Madon, JAR Marques, YA Park (Vice-Chair), WR Roest (Vice-Chair), and S Uścinowicz. See Statement by the Chairman of the Commission on the progress of work in the Commission (CLCS/93) of 18 April 2016, paras 66–68.

26 See art 2(1) of Annex II to UNCLOS.

27 Excerpt of art 3(1)(a) of Annex II to UNCLOS.

28 Expression used in Rule 47(2) of the Rules of Procedure relating to the claimed seaward extent of entitlement.

29 ICJ, Aegean Sea Continental Shelf (Greece v Turkey), Judgment on Jurisdiction of 19 December 1978, ICJ Rep 1980, at 36, para 86.

30 Excerpt from art 3(1)(a) of Annex II to UNCLOS.

31 D Nelson argues that ‘one of the cardinal functions of the Commission must necessarily be to interpret or apply the relevant provisions of the Convention – an essentially legal task’; The Continental Shelf: Interplay of Law and Science’ in Ando, N et al. (eds), Liber Amicorum Judge Shigeru Oda (Kluwer Law 2002) 1235, 1241Google Scholar. R Wolfrum argues also that ‘a competence not referred to in the (Convention) which, nevertheless, is being fulfilled by the Commission is the interpretation, or at least giving guidance, to the interpretation of Article 76 of the Convention’; The Role of International Dispute Settlement Institutions in the Delimitation of the Outer Continental Shelf’ in Lagoni, R and Vignes, D (eds), Maritime Delimitation (Martinus Nijhoff Publishers 2006) 21, 24CrossRefGoogle Scholar.

32 CLCS/11, Scientific & Technical Guidelines of the Commission on the Limits of the Continental Shelf, adopted on 13 May 1999.

33 ibid, point 1.3 (emphasis added).

34 ibid.

35 ibid.

36 ibid.

37 ibid.

38 PCIJ, Interpretation of Judgments Nos 7 and 8 (The Chorzów Factory), Judgment of 16 December 1927, Ser A, No 13, 10.

39 Point 7.3.1 of the Guidelines sets out considerations that are taken into account in the classification of a seafloor high constituting a submarine elevation that is a natural component of the continental margin under para 6 of art 76.

40 Point 7.2 of the Guidelines sets out a non-exhaustive list of criteria that is meant to guide States to distinguish between ‘submarine ridges’ under para 6 of art 76 and ‘oceanic ridges’ under para 3 of art 76.

41 Point 1.3 of the Guidelines.

42 ibid, point 1.3.

43 Art 77(3) of UNCLOS.

44 Excerpt from art 4 of Annex II to UNCLOS.

45 ITLOS, Bay of Bengal, para 407.

46 ibid, para 408.

47 L Lucchini argues that although it is only the coastal State that establishes the outer limits of the continental shelf ‘il doit, en revanche, le faire sur la base des recommandations émises par la Commission’; La delimitation des Frontiéres Maritimes dans la Jurisprudence Internationale: Vue d́Ensemble’ in Lagoni, R and Vignes, D (eds), Maritime Delimitation (Koninklijke Brill 2006) 3, 15CrossRefGoogle Scholar. See also Jarmache, E for a similar opinion; ‘À propos de la Commission’ (2006) 11 Annuaire du droit de la mer 67Google Scholar.

48 In the Bay of Bengal, ITLOS held that ‘the opposability with regard to other States of the limits thus established depends upon satisfaction of the requirements specified in article 76, in particular compliance by the coastal State with the obligation to submit to the Commission information on the limits of the continental shelf beyond 200 [M] and issuance by the Commission of relevant recommendations in this regard. It is only after the limits are established by the coastal State on the basis of the recommendations of the [CLCS] that these limits become “final and binding”.’ ITLOS, Bay of Bengal, para 408. In a note prepared by the DOALOS Secretariat, and which was transmitted to the consideration of States Parties to UNCLOS, during their Meeting of States Parties, it is correctly observed, that ‘[f]or a State to include in its national legislation the general phrase that its continental shelf extends to the outer edge of the continental margin might sound legally correct, but it does not locate the exact position of that outer edge which would be internationally recognized only when considered and recommended by the Commission, accepted by the State, and then incorporated into its national legislation’. SPLOS/64, Issues with respect to article 4 of Annex II to the United Nations Convention on the Law of the Sea – Background paper prepared by the Secretariat (1 May 2001) para 44 (emphasis added). See also Judge Rüdiger Wolfrum, who has described the function of the CLCS as one akin to a legitimator; The Delimitation of the Outer Continental Shelf: Procedural Considerations’ in Badinter, R and Cot, JP (eds), Liber Amicorum Jean-Pierre Cot – Le procès international (Bruylant 2009) 249, 251Google Scholar. In an article from 1989, similar views were expressed by Tullio Treves: ‘il est difficile d'imaginer comment les critères des paragraphes de l'article 76 qui suivent le premier pourraient être appliqués de manière opposable aux autres Etats … sans la sanction d'un organisme technique international indépendant, tel que la Commission des limites du plateau continental’; La limite extérieure du plateau continental: Évolution récente de la pratique’ (1989) 35 AFDI 725, 734Google Scholar.

49 According to art 76(8) of UNCLOS, the outer limits of the continental that are established ‘on the basis of these recommendations shall be final and binding’. B Oxman argues that by virtue of the last sentence in art 76(8) of the Convention, submitting coastal States have been granted ‘an extraordinary power nowhere reproduced with respect to any other maritime limit … They may not be contested.’; The Third United Nations Convention on the Law of the Sea: The Ninth Session (1980)’ (1981) 75 AJIL 221, 230Google Scholar.

50 Oxman argues that a submitting coastal State ‘is not denied the right to reject the recommendations of the Commission’; see (n 49) 230.

51 Art 8 of Annex II to UNCLOS provides that ‘[i]n the case of disagreement by the coastal State with the recommendations of the Commission, the coastal State shall, within a reasonable time, make a revised or new submission to the Commission’. Yet, it has been observed elsewhere that ‘[a] continental shelf boundary that has not been established on the basis of the recommendations of the Commission may still become final and binding in the sense of Article 76(8), depending on the further actions of the coastal State and other States’. First Report of the International Law Association Committee established to study the outer continental shelf (2004) 23 fn 111.

52 Summary of Recommendations of 19 April 2012 to Japan with regard to its submission of 12 November 2008, para 135 <http://www.un.org/depts/los/clcs_new/submissions_files/jpn08/com_sumrec_jpn_fin.pdf>.

53 AM Mantuano notes that ‘[p]arler de jurisprudence n'est donc pas approprié dans ce contexte puisque la Commission n'est pas un organe établi pour juger mais pour analyser et évaluer les données présentées par l'Etat demandeur afin de s'assurer qu'elles sont conformes aux critères contenus à l´Article 76’; Les travaux de la Commission des limites du plateau continental’ in Le plateau continental étendu aux termes de la Convention des Nations Unies sur le droit de la Mer du 10 décembre 1982: optimisation de la demande (Pedone 2004) 399, 403Google Scholar.

54 On this issue see Voelcker, M, ‘Qu'est qu'une “dorsale” au sens de l'article 76 de la Convention de 1982 sur le droit de la mer? Quelques remarques et commentaires à propos des revendications sur le plateau continental arctique’ in Raigon, R Casado and Cataldi, G (eds), L´évolution et l’état actuel du droit international de la mer. Mélanges de droit de la mer offerts à Daniel Vignes (Bruylant 2009) 949–78Google Scholar.

55 Excerpt from art 76(3) of UNCLOS.

56 UNTS No 7302, vol. 499, 312–321.

57 Art 1(a) of the Geneva Convention provides: ‘For the purpose of these articles, the term ‘‘continental shelf’’ is used as referring (a) to the seabed and subsoil of the submarine areas adjacent to the coast but outside the area of the territorial sea, to a depth of 200 metres or, beyond that limit, to where the depth of the superjacent waters admits of the exploitation of the natural resources of the said areas.’

58 In his dissenting opinion in the North Sea cases, ad hoc Judge Sørensen held that ‘[t]he legal concept of the continental shelf cannot reasonably be understood, even in its widest connotation, as extending far beyond the geological concept’. Dissenting Opinion of ad hoc Judge M Sørensen, North Sea Continental Shelf cases, Germany / Denmark Germany / The Netherlands, ICJ Rep 1969, at 249.

59 W Friedmann observed that the inclusion of the exploitability criterion in art 1(a) of the Geneva Convention ‘left the limits of national jurisdiction open’; Selden Redivivus – Towards a Partition of the Seas’ (1969) 63 AJIL 753, 759Google Scholar.

60 See Henkin, L, ‘International Law and the ‘‘Interests’’: The Law of the Seabed’ (1969) 63 AJIL 504–10CrossRefGoogle Scholar.

61 Art 76(3) provides: ‘The continental margin comprises the submerged prolongation of the land mass of the coastal State, and consists of the seabed and subsoil of the shelf, the slope and the rise. It does not include the deep ocean floor with its oceanic ridges or the subsoil thereof.’

62 Brekke, H and Symonds, P, ‘The Ridge Provisions of Article 76 of the UN Convention on the Law of the Sea’ in Nordquist, MH (ed), Legal and Scientific Aspects of Continental Shelf Limits (Martinus Nijhoff Publishers 2004) 169Google Scholar.

63 Doc A/CONF.62/WP.10/Rev.1 (emphasis added).

64 For a general view see Nordquist, MH (ed), United Nations Convention on the Law of the Sea 1982: A Commentary, vol II (Martinus Nijhoff Publishers 1989) 867– 70Google Scholar.

65 Doc A/CONF.62/WP.10/REV.3.

66 Art 76(6) of UNCLOS provides: ‘Notwithstanding the provisions of paragraph 5, on submarine ridges, the outer limit of the continental shelf shall not exceed 350 nautical miles from the baselines from which the breadth of the territorial sea is measured. This paragraph does not apply to submarine elevations that are natural components of the continental margin, such as its plateaux, rises, caps, banks and spurs.’

67 Yet, see Brekke and Symonds who advocate that the definition of the continental margin is ‘apparently by a deliberate choice, made with no reference to geological crustal types in the sense of “continental crust” and “oceanic crust”, but with reference to a geologically unspecified “land mass”’. See ‘The Ridge Provisions of Article 76 of the UN Convention on the Law of the Sea’ (n 62) 180.

68 Allott, P, ‘The concept of international law’ (1999) 10 EJIL 31, 43CrossRefGoogle Scholar.

69 Guidelines, point 7.2.6.

70 See M Voelcker, ‘Qu'est qu'une “dorsale” au sens de l'article 76 de la Convention de 1982 sur le droit de la mer? Quelques remarques et commentaires à propos des revendications sur le plateau continental arctique’; see (n 54) 954. See also comments of Singapore criticizing that not only was the new paragraph 5 bis vague, but it was also agreed ‘à huis clos’ to the exclusion of several Participants. Doc Off. vol. XIII, at 12, para 16.

71 See art 32 of the Vienna Convention on the Law of the Treaties. Concluded at Vienna on 23 May 1969, UNTS, vol. 1155, 1-18232.

72 The first sentence of para 6 of art 76 provides: ‘Notwithstanding the provisions of paragraph 5, on submarine ridges, the outer limit of the continental shelf shall not exceed 350 [M] from the baselines from which the breadth of the territorial sea is measured.’

73 PCIJ, Nationality Decrees Issued in Tunis and Morocco, Advisory Opinion of 7 February 1923, Ser. B 4, at 25.

74 Excerpt from art 76(3) of UNCLOS.

75 ibid.

76 art 76(3) of UNCLOS defines the continental margin as the submerged prolongation of the land mass of the coastal State. This paragraph operates notwithstanding whether the crustal structure of the seafloor high is of oceanic or continental structure. In point 7.2.9 of the Guidelines the CLCS has noted that para 3 of art 76 of UNCLOS is based on a principle of crustal neutrality: ‘The terms “land mass” and “land territory” are both neutral terms with regard to crustal types in the geological sense. Therefore, the Commission feels that geological crust types cannot be the sole qualifier in the classification of ridges and elevations of the sea floor into the legal categories of paragraph 6 of article 76, even in the case of island States.’

77 ITLOS, Bay of Bengal, para 438.

78 In point 7.2.6 of the Guidelines, the CLCS notes that it ‘feels that the provisions of paragraphs 3 and 6 may create some difficulties in defining ridges for which the criterion of 350 M in paragraph 6 may apply on the basis of the origin of the ridges and their composition’.

79 Excerpt from art 76(4)(b) of UNCLOS.

80 Excerpt from point 5.1.3 of the Guidelines.

81 Excerpt from art 76(4)(b) of UNCLOS.

82 Excerpt from point 5.4.6 of the Guidelines.

83 Geology appears to be attributed a subsidiary role where the geomorphological and morphological analysis used to identify a continuous base of slope region does not provide a seafloor high that can be considered an integral part of the continental margin. In its recommendations to Japan on its submission of 12 November 2008, the CLCS noted that the morphological continuity around the southern tip of the Oki-Daito Rise was too tenuous to be considered sufficient for this seafloor high to be considered a part of the submerged prolongation of the land mass of Japan. Rather than refusing to admit its inclusion in the continental margin on the above-mentioned grounds, the CLCS accepted that geological evidence could be dispositive: ‘In order for the Subcommission to consider that a feature with such a tenuous morphological continuity across a saddle as in the case of the “southern tip of the Oki-Daito Rise” would represent part of the submerged prolongation of the mass of a State, the continuity would have to be supported by the existence of geological continuity.’ Summary of Recommendations of CLCS to Japan, para 135.

84 Brekke and Symonds (n 62) 182.

85 Excerpt from art 76(4)(b) of UNCLOS.

86 The Guidelines provide that the CLCS ‘interprets this provision as an opportunity for coastal States to use the best geological and geophysical evidence available to them to locate the foot of the continental slope at its base when the geomorphological evidence given by the maximum change in the gradient as a general rule does not or can not locate reliably the foot of the continental slope’. Excerpt from point 6.1.10 of the Guidelines.

87 The Guidelines provide in point 6.2.6 three examples: (a) convergent (active) continental margins; (b) rifted (non-volcanic) and sheared continental margins; and (c) rifted volcanic continental margins, to which guidance is provided for identifying alternative foot of slope points.

88 Excerpt from point 6.3.11 of the Guidelines.

89 The Guidelines provide in points 6.3.10 and 6.3.11 that in so far concerns (i) rifted (non-volcanic) and sheared continental margins, and (ii) rifted volcanic continental margins the equivalent FOS can be located at the landward limit of the COT. The Guidelines provide in point 6.3.7 that in so far concerns convergent active margins, the equivalent FOS can with ‘acceptable accuracy’ be determined at the ‘seaward limit of the plate boundary’.

90 CLCS/11, point 6.1.9.

91 Point 2.2.8 of the Guidelines provides that ‘[t]he formulation of the test of appurtenance can be described as follows: If either the line delineated at a distance of 60 [nm] from the foot of the continental slope, or the line delineated at a distance where the thickness of sedimentary rocks is at least 1 per cent of the shortest distance from such point to the foot of slope, or both, extend beyond 200 [nm] from the baselines from which the breadth of the territorial sea is measured, then a coastal State is entitled to delineate the outer limits of the continental shelf as prescribed by the provisions contained in article 76, paragraphs 4 to 10.’

92 Summary of Recommendations of the Commission on the Limits of the Continental Shelf in regard to the submission made by Barbados on 8 May 2008, adopted by CLCS on 15 April 2010, para 11; see <http://www.un.org/depts/los/clcs_new/submissions_files/brb08/brb08_summary_recommendations.pdf >.

93 Recommendations of the CLCS to New Zealand with regard to its partial submission of 19 April 2006, adopted on 22 August 2008, para 138; see <http://www.un.org/depts/los/clcs_new/submissions_files/nzl06/nzl_summary_of_recommendations.pdf>.

94 Summary of Recommendations of CLCS to Japan, para 77 (emphasis added).

95 Summary of Recommendations of CLCS to Cook Islands, para 53.

96 ibid para 58.

97 ibid para 80 (emphasis added).

98 ITLOS, Bay of Bengal, para 435.

99 Recommendations of the Commission on the Limits of the Continental Shelf to the Cook Islands, para 58.

100 ITLOS, Bay of Bengal, para 438.

101 ICJ, Case concerning the Continental Shelf (Tunisia / Libyan Arab Jamahiriya), Judgment of 24 February 1982, ICJ Rep 1982, at 54, para 61.

102 Canada noted that it was ‘not in a position to determine whether it agrees with the Russian Federation's Arctic continental shelf submission without the provision of further supporting data to analyse and that Canada's inability to comment at this point should not be interpreted as either agreement or acquiescence by Canada to the Russian Federation's submission’. Note Verbale of Canada of 24 January 2002 to United Nation's Secretary-General, available on the website of DOALOS. In the same vein, Denmark noted that it was ‘not able to form an opinion on the Russian submission. A qualified assessment would require more specific data. Such absence of opinion at this moment does not imply Denmark's agreement or acquiescence to the Russian Federation's submission.’ Note Verbale of Denmark of 5 February 2002 to the United Nations Secretary-General, available on the website of DOALOS. On the reactions of the neighbouring States see Comba, D, ‘The Polar Continental Shelf Challenge: Claims and Exploration of Mineral Sea Resources – An Antarctic and Arctic Comparative Analysis’ (2009) 20 YIEL 158–87Google Scholar.

104 ibid 2.

105 Point 2.2.6 of the Guidelines provides that the CLCS ‘shall use at all times: the provisions contained in paragraph 4(a)(i) and (ii), defined as the formulae lines, and paragraph 4(b), to determine whether a coastal State is entitled to delineate the outer limits of the continental shelf beyond 200 [M]’.

106 United States diplomatic note, at 2.

107 ibid.

108 Recommendations of the Commission on the Limits of the Continental Shelf to the Kingdom of Denmark with regard to the Partial Submission relating to the Northern Continental Shelf of the Faroe Islands, adopted on 24 March 2014, para 34; <http://www.un.org/depts/los/clcs_new/submissions_files/dnk28_09/2014_03_14_SCDNK_REC_COM_20140521.pdf> (emphasis added).

109 ibid.

110 United States diplomatic note (n 89) 3.

111 Summary of Recommendations of the Commission on the Limits of the Continental Shelf in regard to the Submission made by the United Kingdom of Great Britain and Northern Ireland in respect of Ascension Island on 9 May 2008, adopted by the CLCS on 15 April 2010, para 27 (original emphasis) <http://www.un.org/depts/los/clcs_new/submissions_files/gbr08/gbr_asc_isl_rec_summ.pdf>.

112 ibid para 75.

113 Point 7.2.8 of the Guidelines provides: ‘Some ridges (including active spreading ridges) may have islands on them. In such cases it would be difficult to consider that those parts of the ridge belong to the deep ocean floor.’

114 Summary of Recommendations of the CLCS to the United Kingdom in respect of Ascension Island, para 21(ii).

115 ibid para 21(i) (emphasis added).

116 ibid para 43.

117 ibid para 44.

118 ibid.

119 ibid paras 37–38.

120 Summary of Recommendations of the CLCS to the United Kingdom in respect of Ascension Island, para 23(iii).

121 CLCS states also that ‘[t]he United Kingdom regards the rift valley of the spreading axis and the deeps of associated fracture zones as parts of the continental slope of Ascension Island. However, in the view of the Commission, ocean spreading structures, which are normally part of the deep ocean floor, can only form the continental slopes of island landmasses in cases where such structures form part of the discrete seafloor highs from which the island edifices rise. This is not the case for Ascension Island, as its edifice is not morphologically connected to any such discrete seafloor high.’ ibid para 45 (emphasis added).

122 Partial Submission of the Government of the Kingdom of Denmark together with the Government of Greenland, submitted to the CLCS on 16 December 2014.

123 Excerpt from art 76(3) of UNCLOS. J Gao has argued that the ‘[m]id-ocean ridge is the best example of the oceanic ridges’; Continental Shelf beyond 200 Nautical Miles in the Arctic Basin’ (2011) 45 Revue Juridique Thémis 722, 730Google Scholar.

124 H Brekke, ‘The limits of the continental shelf in the Arctic Ocean’, The Norwegian Scientific Academy for Polar Research, Newsletter, No 12 (2014) <http://polar-academy.com/documents/Newsletter_12-June2014.pdf> 3. Yet, the same author has said elsewhere that ‘any morphological seafloor feature around which it is possible to draw a foot of the continental slope, and which is continuous with the foot of the continental slope of the rest of the continental margin, is an integral part of the continental margin under paragraph 4. Therefore, such seafloor features contribute to the outer edge of the continental margin since their foot of the continental slope is eligible to generate an outer edge of margin in accordance with paragraph 4(a)’; see Brekke and Symonds, ‘The Ridge Provisions of art 76 of the UN Convention on the Law of the Sea’ (n 62) 183.

125 Gao argues that ‘consensus has also been reached over the legal status of the Gakkel Ridge. The ridge, also known as the Arctic Mid-Ocean Ridge, is a currently active seafloor spreading system. According to paragraph 3 of article 76, the continental margin ‘‘does not include the deep ocean floor with its oceanic ridges or the subsoil thereof”. Mid-ocean ridge is the best example of the oceanic ridges.’ Gao (n 123) 730.

126 Elferink, AO, ‘The Outer Continental Shelf in the Arctic: The Application of art 76 in the LOS Convention in a Regional Context’ in Elferink, AO and Rothwell, D (eds), The Law of the Sea and Polar Maritime Delimitation and Jurisdiction (Martinus Nijhoff Publishers 2001) 138, 155Google Scholar.

127 Excerpt from art 76(3) of UNCLOS.

128 The Reykjanes Ridge is an active oceanic spreading ridge. It was considered an integral part of the continental margin of Iceland. By contrast to Iceland, whose land mass is composed of oceanic crust, the land mass of Greenland is composed of continental crust. This difference makes it to some extent inappropriate to refer by way of analogy to the situation of the Reykjanes Ridge, when assessing whether the Gakkel Ridge, an active oceanic spreading ridge, is capable of constituting an integral part of the continental margin of any coastal State whose land mass is composed of continental crust.

129 Guidelines, point 7.2.8.

130 Recommendations of the CLCS to Denmark with regard to the partial submission relating to the Northern Continental Shelf of the Faroe Islands, para 27.

131 The geomorphological analysis to identify the base of the slope region under art 76(4)(b) as further developed in point 5.2.1 of the Guidelines has ‘the character of a general rule’. Excerpt from point 5.1.3 of the Guidelines.

132 In its recent recommendations to Iceland the CLCS noted that ‘the base and foot of the continental slope in the Ægir Basin area are unambiguously identifiable on a morphological basis [and therefore] fulfill the requirements of article 76 and Chapter 5 of the Guidelines. The Commission recommends that these FOS points form the basis for the establishment of the outer edge of the continental margin in the Ægir Basin area.’ Summary of Recommendations of the Commission to Iceland, paras 30–31.

133 The relevant part of point 5.2.1 of the Guidelines provides: ‘Bathymetric and geological data provide the evidence to be used in the geomorphological analysis conducted to identify the region defined as the base of the continental slope.’

134 Summary of Recommendations of the CLCS to the United Kingdom in respect of Ascension Island, para 23(iii).

135 According to point 5.4.6 of the Guidelines, the geomorphological analysis can be conducted pursuant to morphology and bathymetry only, where the base of the continental slope can be ‘clearly determined’ on such evidence alone.

136 The first sentence of paragraph 6 of art 76 reads: ‘Notwithstanding the provisions of paragraph 5, on submarine ridges, the outer limit of the continental shelf shall not exceed 350 nautical miles from the baselines from which the breadth of the territorial sea is measure.’

137 See Voelcker, ‘Qu'est qu'une “dorsale” au sens de l'article 76 de la Convention de 1982 sur le droit de la mer? Quelques remarques et commentaires à propos des revendications sur le plateau continental arctique’ (n 54) 955– 65.

138 Excerpt from art 76(6) of UNCLOS.

139 In its recommendations to Iceland regarding the Ægir Basin, the CLCS held in this regard that ‘[t]he depth constraint line lies entirely landward of the distance constraint line in the Ægir Basin. Consequently, the 350 M distance line is the applicable constraint’. Summary of Recommendations of the Commission on the Limits of the Continental Shelf of 10 March 2016, in regard to the Submission made by Iceland in the Ægir Basin Area and in the Western and Southern Parts of Reykjanes Ridge on 29 April 2009, para 41. <http://www.un.org/depts/los/clcs_new/submissions_files/isl27_09/2016_03_10_sc_isl.pdf>.

140 The second sentence of art 76(6) provides: ‘This paragraph does not apply to submarine elevations that are natural components of the continental margin, such as its plateaux, rises, caps, banks and spurs.’

141 Training Manual, VII-33.

142 Denmark is quoted for stating that a submarine elevation that is a natural component of the continental margin should ‘belong to fundamentally the same geological structure as the land territory of the coastal State in question and would support paragraph 5 bis only if that interpretation applied’. Doc Off. vol. XIII, at 17, para 96.

143 NSM Antunes and F Pimentel, ‘Reflecting on the Legal and Technical Interface of Article 76 of the LOSC: Tentative Thoughts on Practical Implementation’, presentation at Conference organized by the ABLOS Advisory Board (28–30 October 2003) <https://www.iho.int/mtg_docs/com_wg/ABLOS/ABLOS_Conf3/PAPER3-1.PDF> 22.

144 Training Manual, VII-32.

145 Brekke and Symonds, ‘The Ridge Provisions of Article 76 of the UN Convention on the Law of the Sea’ (n 62) 187.

146 Gao, J, ‘The Seafloor Highs Issue in Article 76 of the LOS Convention: Some Views from the Perspective of Legal Interpretation’ (2012) 43 ODIL 119, 129CrossRefGoogle Scholar. Oxman notes also that during the seventh session of the Third Conference, there were serious attempts to block any continental shelf entitlement from extending beyond the 200 nm distance line. Others, among which the Soviet Union was a principal figure, were for obvious reasons opposed to limiting the continental shelf entitlement to 200 nm from the baselines, as this formula ‘ignores the geological basis of the continental shelf doctrine … and it might stimulate demands for a universal 300-mile zone, irrespective of geology’. Oxman, B, ‘The Third United Nations Conference on the Law of the Sea: The Seventh Session (1978)’ (1979) 73 AJIL 1, 21CrossRefGoogle Scholar.

147 The International Law Association Committee established to analyse art 76 held in this regard that the qualification in art 76(6) indicates that submarine elevations that are natural components of the continental margin ‘can be distinguished as separate features but at the same time are closely linked to the continental margin. This is the case for features which, although at some point in time were not a part of the continental margin or have become detached from the continental margin, have, through geological processes, become or remained so closely linked to the continental margin as to become or remain a part of it.’ International Law Association, Toronto Conference (2006), Legal Issues of the Outer Continental Shelf, Second Report, at 6. Symonds et al. note that ‘[t]he use of the term natural components in article 76.6 suggests that the features must be physically part of the margin and may be taken to imply a geomorphic and/or geologic definition of what is a natural component.’ Symonds, PA et al. , ‘Ridge Issues’ in Cook, PJ and Carleton, CM (eds), Continental Shelf Limits: The Scientific and Legal Interface (OUP 2000) 300, 301Google Scholar.

148 ITLOS, Bay of Bengal, para 435.

149 For a description of active margins, see point 6.2.3(a) of the Guidelines.

150 For a description of passive margins, see point 6.2.3(b) of the Guidelines.

151 Point 7.3.1 of the Guidelines.

152 ibid.

153 Provisional Scientific and Technical Guidelines of the Commission on the Limits of the Continental Shelf, adopted on 4 September 1998. CLCS/L.6.

154 Art 31(3)(a) of the Vienna Convention provides: ‘There shall be taken into account, together with the context: (a) Any subsequent agreement between the parties regarding the interpretation of the treaty or the application of its provisions.’

155 On this issue see Kunoy, B, ‘The Terms of Reference of the Commission on the Limits of the Continental Shelf: A Creeping Legal Mandate’ (2012) 25 LJIL 109–30CrossRefGoogle Scholar.

156 Guidelines, point 1.4.

157 Art 31(3)(b) of the Vienna Convention provides: ‘There shall be taken into account, together with the context … Any subsequent practice in the application of the treaty which establishes the agreement of the parties regarding its interpretation.’ A Aust notes nevertheless that [g]enerally accepted, however, does not mean that all states parties have to have engaged in a practice, only that all have accepted it, albeit tacitly’; Modern Treaty Law and Practice (CUP 2007) 191Google Scholar.

158 Arbitral Award of 22 December 1963, Interpretation of the air transport services agreement between the United States and France, RIAA, vol. XVI, at 5–74. The Permanent Court stated that ‘[t]he facts subsequent to the conclusion of [a treaty] can only concern the Court in so far as they are calculated to throw light on the intention of the Parties at the time of the conclusion of that Treaty’. PCIJ, Article 3, Paragraph 2 of the Treaty of Lausanne (Frontier between Turkey and Iraq), Advisory Opinion of 21 November 1925, PCIJ Ser. B, No 12, at 24. Yet, Nolte has also observed that the ICJ and arbitral tribunals do ‘not limit its use of subsequent practice to serving as a means of interpretation, but also as a way of recognizing modifications of treaty obligations over time.’ Nolte, G, ‘Subsequent Practice as a Means of Interpretation in the Jurisprudence of the WTO Appellate Body’ in Cannizzaro, E (ed), The Law of Treaties beyond the Vienna Convention (OUP 2011) 138, 141–2Google Scholar.

159 Summary of Recommendations of CLCS to Japan, para 120(v). See also recommendations of the CLCS to Australia in which it refused to admit that the Williams Ridge was a submarine elevation that is a natural component of the continental margin. The CLCS stressed that ‘the data submitted for the WR seems to give only indirect evidence of its nature and origin and the Commission is of the opinion that the geological origin of the WR still remains unresolved. The Commission therefore questions whether the application of paragraph 7.3.1(b) of the Guidelines is justified in the case of the WR. Therefore the Commission does not consider it justified that the WR is regarded a submarine elevation that is a natural component of the continental margin in the sense of article 76, paragraph 6.’ Recommendations of the Commission on the Limits of the Continental Shelf in regard to the Submission made by Australia on 15 November 2004, adopted on 9 April 2008, para 51 <http://www.un.org/depts/los/clcs_new/submissions_files/aus04/Aus_Recommendations_FINAL.pdf>.

160 Recommendation of the CLCS to Australia, para 138. In its recommendations to Norway the CLCS did not admit the contention of Norway that Vøring Spur was a submarine elevation that is a natural component of the continental margin of mainland Norway, which in the view of the CLCS ‘has a different evolution and geological character to the adjacent Vøring Plateau. In the view of the Commission, the Vøring Spur cannot be regarded a submarine elevation that is a natural component of the continental margin of Mainland Norway in the sense of article 76.’ Summary of Recommendations of the CLCS with regard to the partial submission of Norway on 27 November 2006, adopted on 27 March 2009, para 76 <http://www.un.org/depts/los/clcs_new/submissions_files/gbr08/gbr_asc_isl_rec_summ.pdf>.

161 The CLCS held that ‘on balance the JMMC/IP composite high is a submarine elevation that is a natural component of the continental margin of Jan Mayen in the sense of article 76, paragraph 6’. Summary of the Recommendations to Norway, para 77.

162 Summary of the Recommendations to Australia, para 137.

163 ibid (emphasis added).

164 The sub-commission adopted the draft recommendations by majority on 27 February 2014 and then transmitted them to the CLCS on 3 March 2014. CLCS/83, Statement by the Chair on the Progress of Work in the Commission on the Limits of the Continental Shelf (31 March 2014) para 63.

165 Consistent with Rule 35 of the Rules of Procedure of the CLCS, it is obligated to seek to reach consensus in its decision-making. Where this is not possible, ‘decisions of the Commission, subcommission or subsidiary body on all matters of substance shall be taken by a two-thirds majority of the members present and voting’. Excerpt from Rule 37 of the Rules of Procedure of the CLCS.

166 The Under-Secretary-General for Legal Affairs of the United Nations made a statement upon the opening of the fortieth session of the CLCS, inter alia, encouraging the CLCS ‘to make all efforts with a view to finalizing the examination of submissions carried out during the past four years and approving draft recommendations which are currently before the Commission’. CLCS/93, Statement by the Chairman of the Commission on the progress of work in the Commission (18 April 2016) para 5.

167 Summary of Recommendations to Iceland, para 78.

168 Summary of the Recommendations to Australia, para 137.

169 Summary of Recommendations of the CLCS to Iceland, para 78. <http://www.un.org/depts/los/clcs_new/submissions_files/isl27_09/2016_03_10_sc_isl.pdf>.

170 Summary of the Recommendations to Australia, para 137.

171 Summary of Recommendations of the Commission to Iceland, para 76.

172 CLCS/93, Statement by the Chairman of the Commission on the progress of work in the Commission (18 April 2016) para 21.

173 During the Ninth Session at the Third Conference, Iceland is noted for having stated that ‘the new provision regarding submarine ridges meant that the 350-mile limit criterion would apply to ridges which were a prolongation of the land mass of the coastal State concerned.’ Doc Off. vol. XIII, at 17, para 96.

174 It is recalled that by contrast to the 350 M constraint line applicable to submarine ridges, submarine elevations that are natural components of the continental margin are consistent with art 76(6) of UNCLOS subject to the depth constraint.

175 Point 7.1.8 of the Guidelines provides that ‘[t]he distinction between the “submarine elevations” and “submarine ridges” … shall not be based on their geographical denominations and names’.

176 B Oxman, ‘The Third United Nations Conference on the Law of the Sea: The Eighth Session (1978)’ (n 146) 21.

177 It should be added that the statement from 1980 referred to above was made prior to the compromise, which resulted in the inclusion of para 6 of art 76 and, as has been observed elsewhere, seafloor highs which might not be considered part of the continental margin prior to the above-mentioned compromise ‘would be included in the definition of the continental shelf by this amendment.’ International Law Association, Toronto Conference (2006) (n 147) 6.

178 Summary of the Recommendations to the United Kingdom, para 45.

179 International Law Association, Toronto Conference (2006) (n 147) 6.

180 Excerpt from point 7.3.1 of the Guidelines.

181 It should also be noted that in so far so concerns active margins, the Guidelines provide that ‘any crustal fragment or sedimentary wedge that is accreted to the continental margin should be regarded as a natural component of that continental margin.’ Excerpt from point 7.3.1(a) of the Guidelines (emphasis added).

182 The CLCS held that it agreed with the submitting coastal State ‘that the Kermadec and the Colville Ridge system as well as the Three Kings Ridge with the Fantail terrace are natural components of the continental margin’. Summary of the Recommendations of the CLCS to the partial submission of New Zealand, para 145.

183 The entitlement claim of Denmark/Greenland extends along the flanks of the Lomonosov Ridge up to the 200 M distance line from Russia. It results in an outer continental shelf entitlement claim extending up to approximately 950 M from the baselines of Greenland.

184 Executive Summary of the revised partial submission of Russia relating to the Arctic area of 3 (August 2015) at 12, available on the website of DOALOS (emphasis added).

185 Excerpt from diplomatic note of the United States, 3.

186 The International Hydrographic Organization has characterized ridges as ‘elongated narrow elevations of varying complexity having steep sides’ (4th edn, ‘Standardization of Undersea Feature Names’, Bathymetric Publication No 6, November 2008). The CLCS made reference to this definition its its recommendations to the United Kingdom in regard to the Ascension Island (para 26).

187 It should nevertheless be observed that in its note of 30 October 2015 in relation to the transmission to the CLCS of the revised submission of Russia of 3 August 2015, relating to the Arctic, the United States does not reiterate any of the substantive views that were expressed in its note of 18 March 2002, upon the transmission of the original submission of Russia to the CLCS.

188 Summary of the Recommendations of the CLCS to New Zealand, para 136 (emphasis added).

189 In its Note Verbale to the Secretary-General of the United Nations of 11 January 2011, the United Kingdom stated in relation to the CLCS recommendations on Ascension Island that it ‘will await with interest the outcomes of future submissions which raise similar issues of legal interpretation of the Convention, and in particular those submissions which relate to the entitlement of coastal states to continental shelf areas beyond 200 [M] on the basis of mid-ocean ridges’. <http://www.un.org/depts/los/clcs_new/submissions_files/gbr08/gbr_nv_11jan2011.pdf>.

190 ibid.

191 Executive Summary of the Partial Submission of Denmark/Greenland regarding the Northern Continental Shelf of Greenland, at 14 <http://www.un.org/depts/los/clcs_new/submissions_files/dnk76_14/dnk2014_es.pdf>.

192 Summary of the Recommendations to Australia, para 137.

193 Excerpt from point 7.1.6 of the Guidelines.

194 Summary of the Recommendations to Australia, para 137 (emphasis added).

195 Summary of Recommendations of the Commission on the Limits of the Continental Shelf in regard to the submission made by Frane in respect of French Guiana and New Caledonia Region of 22 May 2007, adopted on 2 September 2009, para 48.

196 Russian scientists have expressed support for the conclusion that the Alpha-Mendeleev Ridge is of continental origin. It is argued that the collected data for the purpose of the preparation of the revised partial submission of Russia provides ‘evidence [which] favour[s] the conclusion that the Mendeleev Ridge is composed, at least in part, of attenuated underplated continental crust’. NN Lebedeva-Ivanova et al., ‘Seismic profiling across the Mendeleev Ridge at 82N: Evidence of continental crust’ (2006) Geophysical Journal International 539.

197 Executive Summary of the revised partial submission of Russia relating to the Arctic area of 3 August 2015, at 12.

198 ibid.

199 ibid 13.

200 Russia states in its Executive Summary to the revised submission of 3 August 2015 that ‘[i]n 2005–2014, the Russian organizations carried out a wide range of geological and geophysical studies in order to prepare a partial revised Submission of the Russian Federation in the Arctic Ocean taking into account the recommendations of the Commission of 2002. After 2002, in the central Arctic Basin Russia accomplished: deep seismic sounding of over 4,000 km; over 23,000 km of MCS lines; over 35,000 km of bathymetry survey; 120 stations of geological sampling.’ ibid.

201 Executive Summary of the Partial Submission of Denmark/Greenland regarding the Northern Continental Shelf of Greenland, at 14. It is further provided that evidence suggests ‘that at least parts of the southern Alpha Ridge include highly attenuated continental crust’ (ibid).

202 АФ Морозов et al., ‘Новые геологические данные, обосновывающие континентальную природу области Центрально-Арктических поднятий’, Журнал Региональная геология и металлогения (2013) No 53, 34 – 55 (Morozov, AF et al. , ‘New Geological Data Are Confirming Continental Origin of the Central Arctic Rises’ (2013) 53 Journal Regional Geology and Metallogeny 3455Google Scholar.

203 It is observed that Morozov et al. ‘suggest the Mendeleev Ridge is composed of “thinned underplated continental crust or thickened oceanic crust”, but they prefer a continental origin. In a benchmark paper, Christensen & Mooney (1995) compile global results for the velocity structure of continental crust. Rifted continental crust reveals the highest velocity gradients of the various tectonic environments presented. Velocity gradients for rifted continental crust are considerably less than those observed by Ivanova et al. (2006), potentially arguing against a continental origin for the Mendeleev Ridge. The velocity structure presented by Lebedeva-Ivanova et al. (2006) is more consistent with thickened oceanic crust observed at oceanic plateaus … or volcanic continental margin crust.’ D Dove et al., ‘Bathymetry, controlled source seismic and gravity observations of the Mendeleev ridge; implications for ridge structure, origin, and regional tectonics’ (2010) Geophysical Journal International 494–5.

204 ITLOS, Bay of Bengal, para 438.

205 In the Statement by the Chair of the Commission on the progress of work in the Commission at the Fortieth Session, of 16 April 2016, it is provided that the consideration of eight submissions that remained next in line for consideration, as queued in the order in which they had been received, was deferred: ‘Noting the absence of new communications from States, which indicated developments that would have allowed for the consideration of those submissions, the Commission decided to defer further the establishment of a subcommission to examine any of the above-mentioned submissions.’ CLCS/93, para 77.

206 ITLOS, Bay of Bengal, para. 438.